Anda di halaman 1dari 462

ISSN 1520-295X

Air-Blast Effects on Civil Structures

by
Jinwon Shin, Andrew S. Whittaker, Amjad J. Aref
and David Cormie

Technical Report MCEER-14-0006


October 30, 2014

This research was conducted at the University at Buffalo, State University of New York, and was
supported by MCEER Thrust Area 3, Innovative Technologies.

NOTICE
This report was prepared by the University at Buffalo, State University of New
York, as a result of research sponsored by MCEER. Neither MCEER, associates
of MCEER, its sponsors, the University at Buffalo, State University of New
York, nor any person acting on their behalf:
a.

makes any warranty, express or implied, with respect to the use of any
information, apparatus, method, or process disclosed in this report or that
such use may not infringe upon privately owned rights; or

b.

assumes any liabilities of whatsoever kind with respect to the use of, or the
damage resulting from the use of, any information, apparatus, method, or
process disclosed in this report.

Any opinions, findings, and conclusions or recommendations expressed in this


publication are those of the author(s) and do not necessarily reflect the views
of MCEER or other sponsors.

Air-Blast Effects on Civil Structures


by
Jinwon Shin,1 Andrew S. Whittaker,2 Amjad J. Aref3 and David Cormie4
Publication Date: October 30, 2014
Submittal Date: July 24, 2014
Technical Report MCEER-14-0006

MCEER Thrust Area 3, Innovative Technologies

1
2
3
4

Research Fellow, Department of Civil, Structural and Environmental Engineering, University at


Buffalo, State University of New York
Professor, Chair, and MCEER Director, Department of Civil, Structural and Environmental
Engineering; University at Buffalo, State University of New York
Professor, Department of Civil, Structural and Environmental Engineering, University at Buffalo, State University of New York
Associate Director, Resilience, Security and Risk, Arup, London, United Kingdom

MCEER
University at Buffalo, State University of New York
212 Ketter Hall, Buffalo, NY 14260
E-mail: mceer@buffalo.edu; Website: http://mceer.buffalo.edu

Preface
MCEER is a national center of excellence dedicated to the discovery and development of new knowledge, tools and technologies that equip communities to become more disaster resilient in the face of
earthquakes and other extreme events. MCEER accomplishes this through a system of multidisciplinary,
multi-hazard research, in tandem with complimentary education and outreach initiatives.
Headquartered at the University at Buffalo, The State University of New York, MCEER was originally
established by the National Science Foundation in 1986, as the first National Center for Earthquake
Engineering Research (NCEER). In 1998, it became known as the Multidisciplinary Center for Earthquake Engineering Research (MCEER), from which the current name, MCEER, evolved.
Comprising a consortium of researchers and industry partners from numerous disciplines and institutions
throughout the United States, MCEERs mission has expanded from its original focus on earthquake
engineering to one which addresses the technical and socio-economic impacts of a variety of hazards,
both natural and man-made, on critical infrastructure, facilities, and society.
The Center derives support from several Federal agencies, including the National Science Foundation,
Federal Highway Administration, National Institute of Standards and Technology, Department of
Homeland Security/Federal Emergency Management Agency, and the State of New York, other state
governments, academic institutions, foreign governments and private industry.
The overarching goal of the research described in this report is to characterize the effects of detonations
of high explosives, inside and in the immediate vicinity of the fireball. A computational fluid dynamics
code is verified in 1D and validated in 2D for air blast calculations. The code is used to develop recommendations for modeling the effects of explosions in the near field, to generate new design charts
appropriate for near-field air-blast calculations, and to update design charts for reflection coefficients
for peak overpressure and for scaled impulse as a function of angle of incidence. Recommendations
are provided for calculating values of erosion strain for finite element analysis of reinforced concrete
components subjected to near-field air-blast loadings.

iii

iv

ABSTRACT
The effects of detonations of high explosives are the focus of this report. Analyses are performed
using computational fluid dynamics (CFD) and finite element codes, theoretical formulations and
empirical data.
The effects of detonations of high explosives are characterized in terms of incident and reflected
overpressures and impulses. Calculations are performed to verify and validate a CFD code in 1D
and 2D; estimate blast effects using 1D models; predict incident overpressures and impulses;
provide guidance on the use of reflecting and transmitting boundaries in 2D and 3D models, and
provide recommendations on cell size for CFD analysis. The complex wave field in the Mach
stem region is studied.
Air-blast parameters, including incident and reflected peak overpressures and impulses, and
shock-front arrival times, are typically estimated for protective design using charts developed by
Kingery and Bulmash. The charts underpredict incident and normally reflected peak
overpressures and incident impulse near the face of the charge. Numerical analyses of
detonations of spherical charges of TNT in free air are performed to understand the shortcomings
of current approaches and to provide data for the development of new equations and design
charts for incident and normally reflected overpressures and impulses and for shock-front arrival
time.
Reflection coefficients are often used to transform incident to reflected peak overpressures for
varying angles of incidence. Values for the reflection coefficient are available in textbooks and
technical manuals but these values vary by document, especially in the region of Mach reflection.
Numerical studies are presented to resolve differences between the documents. The
corresponding reflected scaled impulses are also evaluated. Recommendations for design
practice are provided.
Material erosion is often used for simulations of extreme damage to structural components, and
elements are eroded from a finite element mesh based on user-specified criteria. Single element
simulations of concrete are performed to establish reliable values of concrete erosion strain as a
v

function of strain rate, compressive strength, element size and loading condition. Numerical
simulations of a sample reinforced concrete column subjected to blast loadings are undertaken to
demonstrate the utility of the proposed erosion criteria and to characterize, for a single case, the
importance of concrete compressive strength, transverse reinforcement, and axial load on
estimations of damage.

vi

ACKNOWLEDGMENTS
Financial support for this work was provided by MCEER and the State of New York under
Thrust Area 3, Innovative Technologies. This support is gratefully acknowledged. Any opinions,
findings, conclusions or recommendations expressed in this report are those of the authors and
do not necessarily reflect the views of MCEER, the State of New York, or other sponsors.
The authors would like to express their most sincere gratitude to Professor Gray Dargush of the
University at Buffalo for his careful review of this report.

vii

TABLE OF CONTENTS
SECTION

TITLE

SECTION 1
1.1
1.2
1.3
1.4
1.5

BLAST ANALYSIS OF STRUCTURES .......................................................... 1


Introduction ........................................................................................................... 1
State of Practice ..................................................................................................... 2
Goals of the Report ................................................................................................ 4
Notation and Definitions ....................................................................................... 5
Report Organization .............................................................................................. 8

SECTION 2
2.1
2.2
2.3
2.3.1
2.3.2
2.3.3
2.3.4
2.3.5
2.4
2.4.1
2.4.2
2.4.3
2.4.4
2.4.5
2.4.6
2.5
2.6
2.7
2.7.1
2.7.2
2.7.3
2.7.4

MODELING NEAR-FIELD DETONATIONS OF HIGH EXPLOSIVES . 11


Introduction ......................................................................................................... 11
Modeling Detonations in AUTODYN ................................................................ 12
Verification and Validation of AUTODYN ........................................................ 16
Introduction ......................................................................................................... 16
Verification of AUTODYN in 1D for Near-Field Detonations .......................... 17
Verification of AUTODYN in 1D and 2D for Far-Field Detonations ................ 54
Validation of AUTODYN in 2D for Near-Field Detonations ............................. 62
Validation of AUTODYN in 2D for Far-Field Detonations ............................... 69
One-Dimensional Blast Wave Propagation ......................................................... 72
Introduction ......................................................................................................... 72
Mesh Sensitivity Study ........................................................................................ 72
Temperature in the Vicinity of the Charge Face ................................................. 80
Afterburning ........................................................................................................ 88
Blast Wave Propagation in AUTODYN and Air3D ........................................... 97
Equations of State .............................................................................................. 108
Two-Dimensional Blast Wave Propagation ...................................................... 111
Three-Dimensional Wave Propagation ............................................................. 130
Air-Blast Loadings on Sample Reinforced Concrete Columns ......................... 144
Introduction ....................................................................................................... 144
Square Column, Model VI ................................................................................ 145
Circular Column, Model VII ............................................................................. 156
Influence of Target Shape on Air-Blast Loadings ............................................. 164

SECTION 3

INCIDENT AND NORMALLY REFLECTED OVERPRESSURE AND


IMPULSE FOR DETONATIONS IN FREE AIR ....................................... 173
Blast-Resistant Design....................................................................................... 173
Air-Blast Parameter Studies .............................................................................. 173
Kingery and Pannill, 1964: Hemispherical TNT Surface Bursts ...................... 173

3.1
3.2
3.2.1

PAGE

ix

TABLE OF CONTENTS (CONT'D)


SECTION

TITLE

3.2.2
3.2.3

Kingery, 1966: Hemispherical TNT Surface Bursts ......................................... 174


Kingery and Bulmash, 1984: Spherical Free Air and Hemispherical Surface
Bursts of TNT .................................................................................................... 175
Implementation in UFC 3-340-02 using Design Charts .................................... 177
Studies by Swisdak and Bogosian et al. ............................................................ 177
Evaluation of the KB Charts.............................................................................. 178
Numerical Studies ............................................................................................. 179
Spherical Charges used for Numerical Analysis ............................................... 179
Overpressures, Particle Velocities and Accelerations Near the Charge Face ... 180
Mesh Sensitivity Analysis ................................................................................. 188
Incident Overpressure and Impulse ................................................................... 199
Results and Observations .................................................................................. 199
Cell Size for CFD Analysis ............................................................................... 211
Reflected Overpressure and Impulse ................................................................. 212
Results and Observations .................................................................................. 212
Cell Size for CFD Analysis ............................................................................... 221
Shock Front Arrival Time ................................................................................. 222

3.2.4
3.3
3.4
3.5
3.5.1
3.5.2
3.5.3
3.6
3.6.1
3.6.2
3.7
3.7.1
3.7.2
3.8
SECTION 4
4.1
4.2
4.2.1
4.2.2
4.2.3
4.3
4.3.1
4.3.2
4.3.3
4.3.4
SECTION 5
5.1
5.2
5.3

PAGE

DESIGN CHARTS AND POLYNOMIALS FOR AIR-BLAST


PARAMETERS ............................................................................................... 225
Introduction ....................................................................................................... 225
Accuracy of KB and Numerical Predictions of Blast Parameters ..................... 225
Incident Peak Overpressure and Impulse .......................................................... 225
Normally Reflected Peak Overpressure and Impulse ........................................ 226
Shock Front Arrival Time ................................................................................. 226
Updated Polynomials and Charts for Overpressure, Impulse and Arrival Time226
Charts for Design ............................................................................................... 226
Polynomials ....................................................................................................... 228
Sample Results .................................................................................................. 237
Use of Design Charts and Considerations of Non-Ideal Explosives ................. 237
REFLECTED OVERPRESSURE AND IMPULSE AS A FUNCTION OF
ANGLE OF INCIDENCE .............................................................................. 243
Introduction ....................................................................................................... 243
Regular and Mach Reflections .......................................................................... 250
A Numerical Model for Calculating Reflection Coefficients............................ 253
x

TABLE OF CONTENTS (CONT'D)


SECTION

TITLE

5.3.1
5.3.2
5.4
5.4.1
5.4.2
5.4.3
5.5
5.6
5.7
5.8

Modeling and Domains ..................................................................................... 253


Cell Size............................................................................................................. 255
Numerical Models for Calculating Reflected Impulses .................................... 255
Modeling and Domains ..................................................................................... 255
Cell Size............................................................................................................. 258
Effect of Transmitting Boundary....................................................................... 260
Incident and Normally Reflected Overpressures ............................................... 267
Reflected Overpressure as a Function of Angle of Incidence ........................... 267
Incident and Normally Reflected Impulses ....................................................... 280
Reflected Impulse as a Function of Angle of Incidence.................................... 284

SECTION 6

6.7

MODELING CONCRETE EROSION STRAIN FOR BLAST ANALYSIS


OF STRUCTURAL COMPONENTS ........................................................... 295
Introduction ....................................................................................................... 295
Criteria for Eroding Elements ........................................................................... 297
Wave Passage Effects in Reinforced Concrete Columns .................................. 301
Material Models................................................................................................. 309
Concrete ............................................................................................................. 309
Steel Reinforcement .......................................................................................... 315
Single Element Simulations .............................................................................. 319
Blast Analysis of a Sample Reinforced Concrete Column ................................ 329
Introduction ....................................................................................................... 329
Blast Loading..................................................................................................... 330
Mesh Convergence Study .................................................................................. 331
Values of Concrete Erosion Strain .................................................................... 333
Simulation Results ............................................................................................. 335
Alternate Concrete Material Models, Influence of Concrete Compressive
Strength, Confinement, and Axial Pressure Loading ........................................ 346
Values of Erosion Strain for Near-Field Blast Analysis ................................... 366

SECTION 7
7.1
7.2

SUMMARY AND CONCLUSIONS .............................................................. 367


Summary............................................................................................................ 367
Conclusions ....................................................................................................... 369

SECTION 8

REFERENCES ................................................................................................ 373

6.1
6.2
6.3
6.4
6.4.1
6.4.2
6.5
6.6
6.6.1
6.6.2
6.6.3
6.6.4
6.6.5
6.6.6

PAGE

APPENDIX A SPEED OF SOUND......................................................................................... 385


xi

TABLE OF CONTENTS (CONT'D)


SECTION

TITLE

PAGE

A.1
A.2

Introduction ....................................................................................................... 385


Determination of Speed of Sound in Solids and Fluids .................................... 385

APPENDIX B DETONATION WAVE AND PRESSURE ................................................... 387


B.1
Introduction ....................................................................................................... 387
B.2
Hugoniot Curves and Rayleigh Line for the Detonation Front ......................... 388
APPENDIX C INCIDENT AND REFLECTED OVERPRESSURES VERY CLOSE TO
THE CHARGE FACE .................................................................................... 393
C.1
Introduction ....................................................................................................... 393
C.2
AUTODYN Results........................................................................................... 393
APPENDIX D KINGERY AND BULMASH CHARTS ....................................................... 397
APPENDIX E CSCM MODEL PARAMETERS IN LS-DYNA .......................................... 401
APPENDIX F SINGLE-DEGREE-OF-FREEDOM (SDOF) CALCULATIONS FOR
STRAIN RATE ................................................................................................ 403
F.1
Introduction ....................................................................................................... 403
F.2
SDOF Calculations ............................................................................................ 404

xii

LIST OF FIGURES
FIGURE TITLE
2-1
2-2
2-3
2-4
2-5
2-6
2-7
2-8
2-9
2-10
2-11
2-12
2-13
2-14
2-15
2-16
2-17
2-18
2-19
2-20
2-21
2-22
2-23
2-24
2-25
2-26
2-27
2-28
2-29
2-30

PAGE

JWL and ideal gas EOS for TNT ............................................................................... 15


Hydrodynamic parameters at r = 1.1 for a 20 mm mesh .......................................... 19
Hydrodynamic parameters at r = 2.4 for a 20 mm mesh .......................................... 20
Hydrodynamic parameters at r = 2.6 for a 20 mm mesh .......................................... 21
Hydrodynamic parameters at r = 4.5 for a 20 mm mesh .......................................... 22
Hydrodynamic parameters at r = 8 for a 20 mm mesh ............................................. 23
Hydrodynamic parameters at r = 11.7 for a 20 mm mesh ........................................ 24
Hydrodynamic parameters at r = 18 for a 20 mm mesh ........................................... 25
Hydrodynamic parameters at r = 26 for a 20 mm mesh ........................................... 26
Hydrodynamic parameters at r = 34 for a 20 mm mesh ........................................... 27
Hydrodynamic parameters for a 20 mm mesh in AUTODYN (cont.) ...................... 30
Acceleration histories at discrete points along the 1D expansion ............................. 31
Histories of hydrodynamic parameters at r = 1.1 for a 20 mm mesh ....................... 33
Histories of hydrodynamic parameters at r = 2.4 for a 20 mm mesh ....................... 34
Histories of hydrodynamic parameters at r = 2.6 for a 20 mm mesh ....................... 35
Histories for hydrodynamic parameters at r = 4.5 for a 20 mm mesh ...................... 36
Histories of hydrodynamic parameters at r = 8 for a 20 mm mesh .......................... 37
Histories of hydrodynamic parameters at r = 11.7 for a 20 mm mesh ..................... 38
Histories of hydrodynamic parameters at r = 18 for a 20 mm mesh ........................ 39
Histories of hydrodynamic parameters at r = 26 for a 20 mm mesh ........................ 40
Histories of hydrodynamic parameters at r = 34 for a 20 mm mesh ........................ 41
Mesh convergence study at r = 0.5 ........................................................................... 44
Mesh congervence study at r = 0.9 ........................................................................... 45
Mesh convergence study for r = 1.1 ......................................................................... 46
Mesh convergence study for r = 2.4 ......................................................................... 47
Mesh convergence study for r = 2.6 ......................................................................... 48
Overpressure histories at distance, r, from the center of a spherical charge of 18735
kg of TNT; JWL EOS ................................................................................................ 49
Impulse histories at distance, r, from the center of a spherical charge of 18735 kg of
TNT; JWL EOS ......................................................................................................... 52
1/3
Incident overpressure and impulse histories from AUTODYN; Z = 3, 4 and 5 ft/lb ;
cell sizes = 1, 2 and 3 in ............................................................................................. 58
1/3
Incident peak overpressure and impulse from AUTODYN; Z = 3, 4 and 5 ft/lb ; cell
sizes = 1, 2 and 3 in ................................................................................................... 59
xiii

LIST OF FIGURES (CONT'D)


FIGURE TITLE

PAGE

2-31

Summary of AUTODYN (this study) and LS-DYNA and CTH (Browning et al.)
results for incident peak overpressure ....................................................................... 60

2-32

Summary of AUTODYN (this study) and LS-DYNA and CTH (Browning et al.)
results for incident impulse ........................................................................................ 61

2-33

Incident and normally reflected overpressure and impulse histories calculated using
AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.08 m/kg1/3; (Z = 0.20 ft/lb1/3); r = 1.52
................................................................................................................................... 65
Incident and reflected normally overpressure and impulse histories calculated using
AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.12 m/kg1/3; (Z = 0.30 ft/lb1/3); r = 2.28
................................................................................................................................... 66
Incident and normally reflected overpressure and impulse histories calculated using
AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.16 m/kg1/3; (Z = 0.40 ft/lb1/3); r =3.04 .
................................................................................................................................... 67
Incident and normally reflected overpressure and impulse histories calculated using
AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.20 m/kg1/3; (Z = 0.50 ft/lb1/3); r = 3.80
................................................................................................................................... 68
Comparison between AUTODYN calculations and measurements of Goodman
(1960) and Huffington and Ewing (1985) for normally reflected scaled impulses ... 68
Incident and normally reflected overpressure and impulse histories calculated using
AUTODYN for C4 of 1 kg; Z = 1.2 m/kg1/3 (r = 1.2 m) ........................................... 70
Incident and normally reflected overpressure and impulse histories calculated using
AUTODYN for C4 of 0.5 kg; Z = 1.51 m/kg1/3 (r = 1.2 m) ...................................... 71
Comparison between AUTODYN calculations and Frost et al. measurements (2008)
for normally reflected peak overpressure, Pr, and scaled impulse, Ir/W1/3; Z = 1.2 and
1.51 m/kg1/3 ................................................................................................................ 71
Overpressure histories at distance, r, from the center of a spherical charge of 22.68
kg of TNT; JWL EOS; no afterburning ..................................................................... 74
Overpressure histories at distance, r, from the center of a spherical charge of 22.68
kg of TNT; JWL EOS; no afterburning ..................................................................... 77
Overpressure histories at distance, r, from the center of a spherical charge of 22.68
kg of TNT; JWL EOS; no afterburning (cont.) ......................................................... 80
1/3
Overpressure, density and temperature histories at Z = 0.054 m/kg ; spherical
charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh .............................................. 82
1/3
Overpressure, density and temperature histories at Z = 0.058 m/kg ; spherical
charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh .............................................. 83
1/3
Overpressure, density and temperature histories at Z = 0.071 m/kg ; spherical
charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh .............................................. 84

2-34

2-35

2-36

2-37
2-38
2-39
2-40

2-41
2-42
2-43
2-44
2-45
2-46

xiv

LIST OF FIGURES (CONT'D)


FIGURE TITLE

PAGE
1/3

2-47

Overpressure, density and temperature histories at Z = 0.088 m/kg ; spherical


charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh .............................................. 85

2-48

Overpressure, density and temperature histories at Z = 0.177 m/kg ; spherical


charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh .............................................. 86

2-49

Overpressure, density and temperature histories at Z = 0.353 m/kg ; spherical


charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh .............................................. 87

2-50
2-51
2-52
2-53
2-54

Variation of temperature with radial expansion of detonation product ..................... 88


Input of afterburning energy with time for 22.68 kg of TNT .................................... 90
Weight of TNT for volumes of air to release maximum afterburning energy........... 92
Available volumes of air for radial expansion of detonation products ...................... 92
Overpressure histories at distance, r, from the center of a spherical charge of 22.68
kg of TNT for a 0.5 mm mesh; JWL EOS; 129 MJ afterburning .............................. 93
Impulse histories at distance, r, from the center of a spherical charge of 22.68 kg of
TNT for a 0.5 mm mesh; JWL EOS; 129 MJ afterburning ....................................... 95
Overpressure histories as a function of time after detonation at distance x (= r) from
the center of a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh; JWL EOS;
no afterburning........................................................................................................... 99
Air3D overpressure histories as a function of distance and time after detonation for a
spherical charge of 22.68 kg of TNT for a 0.5 mm mesh; ideal gas EOS ............... 104
Influence of EOS on overpressure histories as a function of distance and time after
detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh ............. 109
Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm cells;
Model I-I .................................................................................................................. 112
AUTODYN analysis results for 1D and 2D (I-I with the left transmitting boundary)
models; spherical charge of 22.68 kg of TNT; 0.5 mm mesh; monitoring location 1;
114
Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm cells;
Model I-II................................................................................................................. 115
Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm cells;
Model I-III and Model I-IV ..................................................................................... 117
Reflected overpressure histories on a vertical reflecting surface; Model II; spherical
charge of 22.68 kg of TNT; JWL EOS; varying mesh size ..................................... 123
Pressure fringes as a function of time; Model I; spherical charge of 22.68 kg of TNT;
JWL EOS; TNT density = 1636.8 kg/m3; 1.0 mm mesh ......................................... 125
Reflected overpressure histories on a vertical reflecting surface; Model III; spherical
charge of 22.68 kg of TNT; JWL EOS; varying mesh size ..................................... 128

2-55
2-56

2-57
2-58
2-59
2-60

2-61
2-62
2-63
2-64
2-65

1/3

1/3

xv

LIST OF FIGURES (CONT'D)


FIGURE TITLE
2-66

2-67
2-68
2-69
2-70
2-71
2-72
2-73
2-74
2-75
2-76
2-77
2-78
2-79
2-80
2-81
2-82
2-83
2-84

PAGE

Reflected overpressure histories on a vertical reflecting surface; Model IV; spherical

charge of 22.68 kg of TNT; JWL EOS; 4 mm mesh (4.027 mm for y = 149 mm);
varying perpendicular dimension, y ......................................................................... 132
Reflected overpressure histories on a vertical reflecting surface; Model II and Model
IV; spherical charge of 22.68 kg of TNT; JWL EOS; y = 100 mm for 3D analysis 135
Reflected overpressure histories on a vertical reflecting surface; Model II and Model
IV; spherical charge of 22.68 kg of TNT; JWL EOS; y = 200 mm for 3D analysis 137
Reflected overpressure histories on a vertical reflecting surface; Model II and Model
IV; spherical charge of 22.68 kg of TNT; JWL EOS; 2 mm mesh ......................... 139
Reflected overpressure histories on a vertical reflecting surface; Model V; spherical
charge of 22.68 kg of TNT; JWL EOS; y = 200 mm ............................................. 141
Reflected overpressure histories on a vertical reflecting surface; Model V; spherical
charge of 22.68 kg of TNT; JWL EOS; 2 mm and 4 mm meshes ........................... 142
Model VI; spherical charge of 22.68 kg of TNT; square column 10001000 mm 146
Monitoring locations for Model VI; spherical charge of 22.68 kg of TNT; square
column 10001000 mm .......................................................................................... 147
Reflected overpressure histories; Model VI; spherical charge of 22.68 kg of TNT;
JWL EOS ................................................................................................................. 148
Net loading and impulse histories; Model VI .......................................................... 153
Comparison of reflected overpressure histories from Models V and VI ................. 154
Model VII; spherical charge of 22.68 kg of TNT; 1000 mm diameter column ...... 157
Monitoring locations for Model VII; spherical charge of 22.68 kg of TNT; 1000 mm
diameter column ...................................................................................................... 158
Reflected overpressure histories; Model VII; spherical charge of 22.68 kg of TNT;
JWL EOS ................................................................................................................. 159
Net load and impulse histories; Model VII; spherical charge of 22.68 kg of TNT;
JWL EOS ................................................................................................................. 163
Reflected overpressure histories at 0 mm above the ground surface; Models VI and
VII; spherical charge of 22.68 kg of TNT; JWL EOS ............................................. 164
Reflected overpressure histories at 149 mm above the ground surface; Models VI
and VII; spherical charge of 22.68 kg of TNT; JWL EOS ...................................... 167
Reflected overpressure histories at 298 mm above the ground surface; Models VI
and VII; spherical charge of 22.68 kg of TNT; JWL EOS ...................................... 169
Reflected overpressure histories at 447 mm above the ground surface; Models VI
and VII; spherical charge of 22.68 kg of TNT; JWL EOS ...................................... 171

xvi

LIST OF FIGURES (CONT'D)


FIGURE TITLE

PAGE

3-1

Incident overpressures and particle velocities and accelerations near the face of the
charge ; TNT weight of 23 kg; 0.1 mm cells ........................................................... 182

3-2

Incident overpressures and particle velocities and accelerations near the face of the
charge ; TNT weight of 23 kg; 0.05 mm cells ......................................................... 183

3-3

Incident overpressures and particle velocities and accelerations near the face of the
charge ; TNT weight of 23 kg; 0.1 and 0.05 mm cells ............................................ 184

3-4

Incident overpressures and particle velocities and accelerations near the face of the
charge; TNT weight of 960 kg; 0.1 mm cells .......................................................... 185

3-5

Incident overpressures and particle velocities and accelerations near the face of the
charge; TNT weight of 960 kg; 0.05 mm cells ........................................................ 186

3-6

Incident overpressures and particle velocities and accelerations near the face of the
charge; TNT weight of 960 kg; 0.1 and 0.05 mm cells ........................................... 187

3-7
3-8

AUTODYN 2D simulation using 1D remapped data for Z = 1.0 m/kg1/3 ............... 194
Ratios of incident peak overpressure for different cell sizes; 0.0553 Z < 40.0
m/kg1/3 ...................................................................................................................... 200
Ratios of incident impulse for different cell sizes; 0.0553 Z < 40.0 m/kg1/3 ....... 201
Ratios of incident peak overpressure and impulse for the fine meshes; charge
weights of 18735, 960 and 23 kg; 0.0553 Z < 40.0 m/kg1/3 ................................. 202
Incident peak overpressure versus scaled distance .................................................. 203
Incident impulse versus scaled distance .................................................................. 207
Ratios of reflected peak overpressure for different cell sizes; 0.0553 Z < 40.0
m/kg1/3 ...................................................................................................................... 213
Ratios of reflected impulse for different cell sizes; 0.0553 Z < 40.0 m/kg1/3 ...... 214
Ratios of reflected peak overpressure and impulse for the smallest cell sizes; charge
weights of 18735, 960 and 23 kg; 0.0553 Z < 40.0 m/kg1/3 ................................. 215
Normally reflected peak overpressure versus scaled distance ................................. 216
Normally reflected impulse versus scaled distance ................................................. 219
AUTODYN calculations and KB predictions for specific arrival time ................... 222
Ratios of AUTODYN calculations and KB predictions of specific arrival time .... 223
Effect of charge weight on specific arrival times calculations: 18735, 960 and 23 kg
charges ..................................................................................................................... 223
Spherical charge and radial expansions, r , of 1.025, 1.050 and 1.100 .................. 227
Incident peak overpressure impulse and arrival time, spherical free-air bursts ...... 233
Reflected peak overpressure and impulse, spherical free-air bursts ........................ 234
Proposed polynomial charts in SI units, spherical free-air bursts ........................... 235

3-9
3-10
3-11
3-12
3-13
3-14
3-15
3-16
3-17
3-18
3-19
3-20
4-1
4-2
4-3
4-4

xvii

LIST OF FIGURES (CONT'D)


FIGURE TITLE
4-5
4-6
5-1
5-2
5-3
5-4
5-5
5-6
5-7
5-8
5-9
5-10
5-11
5-12
5-13
5-14
5-15
5-16
5-17
5-18
5-19
5-20

5-21

PAGE

Proposed polynomial charts in US units, spherical free-air bursts .......................... 236


Use of the proposed polynomials and charts ........................................................... 240
Shock wave reflection phenomena from explosion in free air (adapted from Norris et
al. 1959) ................................................................................................................... 244
Reflection coefficient versus angle of incidence ..................................................... 246
Reflected scaled impulse versus angle of incidence (UFC 3-340-02 (DoD 2008)) 247
Incident, Ps , and normally reflected, Pr , peak overpressures developed by Kingery
and Bulmash (1984) and in Chapter 4 ..................................................................... 248
1/3
1/3
Incident, I s /W , and normally reflected, I r /W , scaled impulses developed by
Kingery and Bulmash (1984) and in Chapter 4 ....................................................... 249
Regular and Mach reflections (e.g., Smith and Hetherington 1994) ....................... 250
Dual solution domain (Ivanov et al. 2001) .............................................................. 251
Angle of incidence for Mach reflection as a function of Mach number for an ideal
gas (Kinney and Graham 1985) ............................................................................... 252
Incident overpressure as a function of Mach number for an ideal gas (Kinney and
Graham 1985) .......................................................................................................... 252
2D numerical model of air for calculating reflection coefficients ........................... 255
2D numerical models of air for calculating reflected impulses ............................... 257
Models 1, 2 and 3 in AUTODYN for angle of incidence of 80 ............................. 258
Mesh sensitivity study for simulating reflected impulse; Model 1.......................... 259
Effect of transmitting boundary on calculation of reflected impulse for Z = 0.16
m/kg1/3 ...................................................................................................................... 261
Effect of transmitting boundary on calculation of reflected impulse for Z = 0.40
m/kg1/3 ...................................................................................................................... 263
Effect of transmitting boundary on calculation of reflected impulse for Z = 1.2
m/kg1/3 ...................................................................................................................... 265
Incident and normally reflected peak overpressures ............................................... 268
Overpressure reflection coefficients as a function of angle of incidence ................ 272
1/3
1/3
Reflected overpressure histories; Z = 0.16, 0.40, 0.80 and 1.2 m/kg (1 m/kg
=2.52 ft/lb1/3; 1 MPa = 145 psi) ............................................................................... 276
Overpressure reflection coefficients with angle of incidence from TM 5-858-3 (DoA
1984), UFC 3-340-02 (DoD 2008), numerical analyses by AUTODYN; Ps in units
of MPa ..................................................................................................................... 278
Incident and normally reflected scaled impulses ..................................................... 281

xviii

LIST OF FIGURES (CONT'D)


FIGURE TITLE
5-22
5-23
5-24
5-25
5-26
5-27
6-1
6-2
6-3
6-4
6-5
6-6
6-7
6-8
6-9
6-10
6-11
6-12
6-13
6-14
6-15
6-16
6-17

PAGE

Reflected scaled impulse as a function of angle of incidence (Z in units of m/kg1/3)....


................................................................................................................................. 285
Reflected overpressure and scaled impulse histories for cell sizes of R/400 and R/800;
= 30, 40, 45, 50 and 55; Z = 0.16 m/kg1/3 ....................................................... 287
Reflected overpressure and scaled impulse histories for cell sizes of R/400 and R/800;
= 30, 40, 45, 50 and 55; Z = 0.40 m/kg1/3 ....................................................... 288
Reflected overpressure and scaled impulse histories for cell sizes of R/400 and R/800;
= 30, 40, 45, 50 and 55; Z = 1.2 m/kg1/3 ......................................................... 289
Pressure fringes at t = 1.0 ms for Z = 0.4 m/kg1/3; R/800 cells ................................ 291
Reflected scaled impulse as a function of angle of incidence; Ps in units of MPa . 292
Reinforced concrete column (units in meters) ......................................................... 302
Rectangular pulse used for the elastic simulation of the sample RC column .......... 303
Cross section at the mid-height of the sample RC column ...................................... 304
Stress histories in the x-direction for elements No. 1, 3 and 7 ................................ 304
Stress histories in the x-direction for elements No. 4, 5, 6 and 7 ............................ 304
Cross section at the mid-height of the 1.2-m-deep column ..................................... 306
Comparison of stress histories of elements No. 1 through 7 in the x-direction
between 0.6-m- and 1.2-m-deep columns (cont.) .................................................... 308
Differences in the stress histories of elements No. 4, 5, 6 and 7 in the x-direction
between the 0.6-m- and 1.2-m-deep columns .......................................................... 308
Dynamic Increase Factor for compressive and tensile strength of concrete ........... 310
General shape of the CSCM yield surface (Murray 2007) ...................................... 310
Stress-strain relationship for Grade 60 rebar at a strain rate of 510-4 s-1 .............. 317
Compressive stressstrain curves for concrete models of the CSCM and by Popovics
(1970) for the compressive strength of 35.5 MPa ................................................... 319
Loading conditions used in single element simulations when the peak tensile load =
10 MPa ..................................................................................................................... 320
Time step convergence analysis for an element size of 404040 mm ................. 322
Erosion strain versus traction rate in single element simulations for different peak
tensile tractions, fc = 35.5 MPa ............................................................................... 323
Erosion strain versus traction rate in single element simulations for different peak
tensile tractions, fc = 50 MPa .................................................................................. 325
Erosion strain versus traction rate in single element simulations for different element
sizes, fc = 35.5 MPa ................................................................................................ 327

xix

LIST OF FIGURES (CONT'D)


FIGURE TITLE

PAGE

6-18

Erosion strain versus traction rate in single element simulations for different element
sizes, fc = 50 MPa ................................................................................................... 328

6-19
6-20
6-21
6-22

Blast loading calculations (units: m) ....................................................................... 329


Reflected pressure histories on the sample reinforced concrete column ................. 332
Mesh convergence study of the sample RC column ................................................ 333
FE models of the sample RC column to estimate traction rates and peak tensile
stresses (tractions) for 30 elements near or on the rear face of the column at its midheight ....................................................................................................................... 334
Maximum principal tensile stress history of an elastic element in models 1 and 2, 40
mm mesh .................................................................................................................. 334
Maximum principal tensile stress histories for 30 elements near or on the rear face of
the column at its mid-height; model 2 ..................................................................... 335
Displacement histories at the mid-height of the column for the 202020 mm mesh
................................................................................................................................. 336
Displacement histories at the mid-height of the column for the 303030 mm mesh .
................................................................................................................................. 336
Displacement histories at the mid-height of the column for the 404040 mm mesh .
................................................................................................................................. 337
Displacement histories at the mid-height of the column for the 606060 mm mesh .
................................................................................................................................. 337
Reaction histories at the bottom of the column for the 202020 mm mesh ........ 337
Reaction histories at the bottom of the column for the 303030 mm mesh ........ 338
Reaction histories at the bottom of the column for the 404040 mm mesh ........ 338
Reaction histories at the bottom of the column for the 606060 mm mesh ........ 338
Internal energy histories for the 202020 mm mesh ........................................... 339
Internal energy histories for the 303030 mm mesh ........................................... 339
Internal energy histories for the 404040 mm mesh ........................................... 340
Internal energy histories for the 606060 mm mesh ........................................... 340
Simulation results for the 202020 mm mesh at time = 25 msec ........................ 342
Simulation results for the 303030 mm mesh at time = 25 msec ........................ 343
Simulation results for the 404040 mm mesh at time = 25 msec ........................ 344
Simulation results for the 606060 mm mesh at time = 25 msec ........................ 345
Maximum principal stress histories for 30 elements near or on the rear face of the
column, near its mid-height, CONCRETE_DAMAGE_REL3 ............................... 347

6-23
6-24
6-25
6-26
6-27
6-28
6-29
6-30
6-31
6-32
6-33
6-34
6-35
6-36
6-37
6-38
6-39
6-40
6-41

xx

LIST OF FIGURES (CONT'D)


FIGURE TITLE
6-42

6-43
6-44
6-45
6-46
6-47
6-48
6-49
6-50
6-51
6-52
6-53
6-54
6-55
6-56
6-57
6-58
B-1
B-2
B-3
C-1
C-2
C-3

PAGE

Internal energy histories for the concrete material models of the CSCM using

calculated erosion strains and an erosion algorithm of damage function, and the
CONCRETE_DAMAGE_REL3 model using calculated erosion strains ............... 348
Simulation resutls for the CONCRETE_DAMAGE_REL3 model......................... 349
Internal energy histories for fc = 50 MPa .............................................................. 351
Simulation results for the 20 mm mesh at time = 25 msec, fc = 50 MPa .............. 352
Simulation results for the 30 mm mesh at time = 25 msec, fc = 50 MPa .............. 352
Simulation results for the 40 mm mesh at time = 25 msec, fc = 50 MPa .............. 353
Loading and boundary conditions of the single element for studying the effects of
confinement ............................................................................................................. 355
Displacement-controlled loading at the top of the single element .......................... 355
Confinement effects on stress-strain relationships, 202020 mm element .......... 356
3D symmetry model of concrete cylinder (units: mm) ............................................ 356
Compressive stress versus strain relationships ........................................................ 357
Simulation results for the 20 mm mesh with the spacing of transverse reinforcement
of 20 cm at time = 25 msec, fc = 35.5 MPa ........................................................... 358
Simulation results for the 30 mm mesh with the spacing of transverse reinforcement
of 20 cm at time = 25 msec, fc = 35.5 MPa ........................................................... 358
Simulation results for the 40 mm mesh with the spacing of transverse reinforcement
of 20 cm at time = 25 msec, fc = 35.5 MPa ........................................................... 359
Simulation results for the 20 mm mesh with axial pressure loading, fc = 35.5 MPa ..
................................................................................................................................. 360
Simulation results for the 30 mm mesh with axial pressure loading, fc = 35.5 MPa ..
................................................................................................................................. 362
Simulation results for the 40 mm mesh with axial pressure loading, fc = 35.5 MPa ..
................................................................................................................................. 364
States across a detonation (shock) wave in unsteady flow (adapted from Fickett and
Davis 1979) .............................................................................................................. 388
States across a detonation (shock) wave in steady flow flow (adapted from Fickett
and Davis 1979) ....................................................................................................... 391
Hugoniot curves and Rayleigh line (adapted from Smith and Hetherington 1994) 392
Incident and reflected peak overpressures near the charge face for cell sizes of 0.1
mm and 0.05 mm ..................................................................................................... 394
Incident peak overpressures inside and reflected peak overpressures at the charge
face as a function of cell size ................................................................................... 394
Reflection coefficients for distances between 151 and 155 mm ............................. 396
xxi

LIST OF FIGURES (CONT'D)


FIGURE TITLE
C-4
D-1
D-2
D-3
D-4
F-1

PAGE

Peak particle velocities for distances between 151 and 155 mm............................. 396
Air-blast parameters as a function of scaled distance in SI units; spherical free-air
burst of TNT ............................................................................................................ 397
Air-blast parameters as a function of scaled distance in US units; spherical free-air
burst of TNT ............................................................................................................ 398
Air-blast parameters as a function of scaled distance in SI units; hemispherical
surface burst of TNT ................................................................................................ 399
Air-blast parameters as a function of scaled distance in US units; hemispherical
surface burst of TNT ................................................................................................ 400
Strain-rate histories for three single element simulations........................................ 405

xxii

LIST OF TABLES
TABLE

TITLE

2-1
2-2
2-3
2-4
2-5
2-6
2-7
2-8
2-9
2-10
2-11
2-12
2-13

JWL EOS parameters for TNT (Dobratz and Crawford 1985) ................................. 13
Monitoring locations used by Browning et al. (2013) ............................................... 55
JWL EOS parameters for modeling TNT used by Browning et al. (2013) ............... 55
Values of input parameters for modeling air for far-field verification ...................... 55
Results for incident peak overpressure; cell size of 1 in; units of psi........................ 56
Results for incident peak overpressure; cell size of 2 in; units of psi........................ 56
Results for incident peak overpressure; cell size of 3 in; units of psi........................ 57
Results for incident impulse; cell size of 1 in; units of psi-ms .................................. 57
Results for incident impulse; cell size of 2 in; units of psi-ms .................................. 57
Results for incident impulse; cell size of 3 in; units of psi-ms .................................. 57
Properties for 50/50 Pentolite and C4 ........................................................................ 63
JWL EOS parameters for 50/50 Pentolite and C4 (Dobratz and Crawford 1985) .... 63
Normally reflected scaled impulse from AUTODYN calculations and measurements
of Goodman (1960) and Huffington and Ewing (1985); units of MPa-ms/kg1/3 ....... 64
Calculations for the afterburning energy at each interval between radial expansions
of detonation products for a spherical charge of 22.68 kg of TNT. .......................... 90
AUTODYN dataset for incident calculations .......................................................... 189
Cell sizes for incident peak overpressure and impulse; 960 kg charge with a radius of
520 mm .................................................................................................................... 191
Cell sizes for incident peak overpressure and impulse; 18735 kg charge with a radius
of 1400 mm .............................................................................................................. 192
Cell sizes for incident peak overpressure and impulse; 23 kg charge with a radius of
150 mm .................................................................................................................... 193
AUTODYN dataset for reflected calculations ......................................................... 196
Cell sizes for reflected peak overpressure and impulse; 960 kg charge with a radius
of 520 mm ................................................................................................................ 197
Cell sizes for reflected peak overpressure and impulse; 18735 kg charge with a
radius of 1400 mm ................................................................................................... 198
Cell sizes for reflected peak overpressure and impulse; 23 kg charge with a radius of
150 mm .................................................................................................................... 198
Constants of polynomials for incident and normally reflected peak overpressure,
impulse and arrival time for 0.0553 Z 40 m/kg1/3 ............................................ 230
Constants of polynomials for incident and normally reflected peak overpressure,
impulse and arrival time for ranges of scaled distance 0.0553 Z < 0.5 m/kg1/3 and
0.5 Z 40 m/kg1/3 ................................................................................................ 231

2-14
3-1
3-2
3-3
3-4
3-5
3-6
3-7
3-8
4-1
4-2

PAGE

xxiii

LIST OF TABLES (CONT'D)


TABLE

TITLE

4-3

Values and slopes at Z = 0.5 m/kg1/3 on the polynomials cures plotted using the two
ranges of scaled distance ......................................................................................... 232
Sample calculations for spherical 25 kg and 1000 kg charges and scaled distances of
0.25 m/kg1/3 and 1.00 m/kg1/3 .................................................................................. 239
Incident and reflected overpressures for normal incidence ..................................... 269
Ratios of AUTODYN calculations to proposed polynomial values and UFC 3-34002 (DoD 2008) predictions ...................................................................................... 270
Reflection coefficients as a function of angle of incidence, , and scaled distance, Z,
calculated using AUTODYN ................................................................................... 271
Incident and normally reflected scaled impulses in SI units of MPa-ms/kg1/3 and US
units of psi-ms/lb1/3 .................................................................................................. 282
Ratios of AUTODYN calculations to polynomial and UFC 3-340-02 predictions for
reflected scaled impulse ........................................................................................... 283
Reflected scaled impulses as a function of angle of incidence, , and scaled distance,
Z, calculated using AUTODYN (units: MPa-ms/kg1/3; 1 MPa-ms/kg1/3 = 111 psims/lb1/3) .................................................................................................................... 286
Assumed model parameters for Grade 6d0 reinforcement (adopted from Brvik et al.
2001) ........................................................................................................................ 316
Gruneisen EOS parameters for steel (Tan et al. 2009) ............................................ 318
Johnson and Cook damage model parameters for Weldox 460 E steel (Brvik et al.
2001) ........................................................................................................................ 318
Influence of boundary conditions in the single element simulations for a compressive
strength of 35.5 MPa at 200 MPa/msec ................................................................... 321
Percentage erosion of the cross section at the mid-height of the column ................ 341
CSCM model parameters in LS-DYNA .................................................................. 401
Properties of single elements ................................................................................... 403

4-4
5-1
5-2
5-3
5-4
5-5
5-6

6-1
6-2
6-3
6-4
6-5
E-1
F-1

PAGE

xxiv

SECTION 1
BLAST ANALYSIS OF STRUCTURES
1.1 Introduction
The effects of air-blast loadings are routinely considered for the design of mission-critical
buildings, bridges and infrastructure. Textbooks and technical manuals for blast-resistant design
are available (e.g., Smith and Hetherington 1994, FEMA 2003, DOD 2008, Krauthammer 2008,
Cormie et al. 2009, Dusenberry 2010), but most of the guidance was written around large farfield loadings. Empirical charts are provided in some of these documents to compute air-blast
loadings (incident and reflected overpressures and impulses) but these charts have not been
validated in the near field because either the target is within the fireball or the overpressures are
too high to be measured by commercially available transducers. Some of the mid-field data used
to generate the charts were inferred and not measured directly.
The empirical charts return values of incident and reflected peak overpressure, incident and
reflected (specific) impulse, arrival time, positive phase duration and shock front velocity.
Overpressure histories are calculated using the Friedlander equation (e.g., Smith and
Hetherington 1994):

t
t
ps (t ) Ps 1 e bt / to and pr (t ) Pr 1 e bt / to
to
to

(1-1)

where ps and pr are incident and reflected overpressures, respectively; Ps and Pr are incident
and reflected peak overpressures, respectively; t is time; to is positive phase duration, and b is a
waveform parameter. The parameter b in Equation 1-1 can be back-calculated using the peak
overpressure, positive phase duration, and the impulse calculated using the charts for a given
scaled distance and charge weight.
Over the past decade, the rapid development of computational fluid dynamics (CFD) codes has
enabled the simulation of complex blast phenomena. CFD codes are deployed in LS-DYNA
(LSTC 2013) and AUTODYN (ANSYS 2009). Air3D (Rose 2006) is a CFD code written

specifically for the calculation of air-blast loadings. These codes, once verified and validated,
enable a critical review of empirical design charts and approaches, with a focus here on the most
common threat, namely, small-to-medium weight weapons and a small stand-off distance: nearfield detonations.

1.2 State of Practice


The practice of blast-resistant design is bimodal; namely, 1) CFD simulations and finite element
analysis, and 2) empirical charts, simplified loading functions, and single degree of freedom
analysis. The former should inform the latter. The latter were developed in the 1950s, with a
focus on design of above-ground military infrastructure against the threat posed by far-field
thermonuclear detonations. The seminal texts of Biggs (1964) and Norris et al. (1959) enabled
design against these threats. Loadings were calculated using first principles calculations of
idealized detonations (e.g., Brode 1955, Henrych 1979, Kinney and Graham 1985) and charts
presented in government manuals (e.g., TM-5-1300 (Department of the Army, Navy and Air
Force 1990) and text books (e.g., Smith and Hetherington 1994, Cormie et al. 2009, Dusenberry
2010).
The blast-resistant design of a structure or a component thereof first involves modeling the
effects of the detonation. Traditional practice in the United States, United Kingdom, Australia
and elsewhere involves the use of empirical charts developed by Kingery and Bulmash (1984),
as presented in US government documents such as UFC 3-340-02 (DOD 2008). Computer codes,
such as CONWEP (Hyde 1992), implement the Kingery and Bulmash (KB) polynomials. The
charts provide estimates of incident and normally reflected overpressures for spherical TNT
explosions in free air and hemispherical TNT explosions on a rigid reflecting surface. Other
charts are available in textbooks and government manuals to estimate reflected overpressures and
impulses for angles of incidence other than 0 degrees (the so-called normal reflection).
For large far-field detonations of high explosives, one pressure history can represent the load
effect over the height and width of a component or structure. The typical history is characterized
by an instantaneous rise in pressure to a peak value and a linear decay to ambient pressure. This
loading history is a simplification of the true loading history that shows an exponential decay in
2

overpressure to ambient (the positive phase) and a period of underpressure, which may be
interrupted by the arrival of a secondary shock front. The Friedlander curve describes the
exponential decay to ambient pressure, noting that values for the parameters that define the curve
must be determined by curve fitting to a dataset.
For near-field detonations, the expansion of the detonation products and afterburning affect the
amplitude and shape of the overpressure histories. Charge shapes different from a sphere or
hemisphere are not addressed in sufficient detail, and the point of detonation within the charge is
assumed to be the middle of the sphere. Alternate trigger points for the detonation, which can
significantly influence the overpressure histories in the near-field (Sherkar 2010), cannot be
considered.
Detonations can be modeled numerically using the finite difference method, which is well suited
for fluid dynamics and can solve the Euler equations (e.g., Rose 2006), but the chemistry of a
detonation is rarely modeled. Many equations of states (EOS) have been developed to model
expanding explosives (e.g., Becker-Kistiakowsky-Wilson (BKW) EOS, Jones-Wilkins-Lee
(JWL) EOS, Kihara-Hikita-Tanaka (KHT) EOS, Lennard-Jones-Devonshire (LJD) EOS). The
EOS defines the relationships between state variables of thermodynamics such as pressure,
volume, density and internal energy. The JWL EOS is the most widely used EOS for explosives
due to its ease of implementation in hydrodynamic calculations. For air, the ideal gas EOS is
typically adopted in the modeling scheme.
Components of structures, and indeed structural systems, can be modeled in great detail (e.g., a
micro-level finite element model) or simplistically using single-degree-of-freedom systems,
wherein the mechanical properties of continuous systems are transformed using shape and
resistance functions for idealized boundary conditions. For detailed analysis, the finite element
(FE) method can be used. Constitutive models, which define the relationships between stress,
strain, strain rate, and temperature have been developed for different materials (e.g., Johnson and
Cook 1983, Malvar and Simons 1996, Murray 2007a) and are deployed at the element (brick)
level for FE analysis. For single-degree-of-freedom analysis, strain-rate effects are addressed
using dynamic increase factors, which have been developed for concrete (e.g., Malvar and Ross

1998, Hao and Zhou 2007) and metals (e.g., Johnson and Cook 1983, DOD 2008), for
compressive, tensile, and shearing forces.
When concrete structures are subjected to close-in detonations of high explosives, material
failure may occur with the associated fragmentation (damage) of concrete. The effect of damage
can be simulated in FE models using erosion algorithms, which eliminate material from a model
to avoid the numerical errors associated with highly distorted FE meshes. Erosion has been
implemented in FE codes (e.g., ABAQUS, AUTODYN, and LS-DYNA) using erosion
parameters or damage algorithms. Erosion cannot be addressed with single-degree-of-freedom
analysis.

1.3 Goals of the Report


The primary goals of this report are to characterize the effects of detonations of high explosives
and their influence on structures, for the purpose of informing blast-resistant design. The specific
objectives are,
1. to verify and validate a CFD code for calculating air blast effects in the near and far fields,
2. to develop validated methodologies and guidance for modeling detonations,
3. to update the Kingery and Bulmash charts and polynomials for incident and normally
reflected overpressures and impulses, and shock front arrival times, with an emphasis on
the near field,
4. to reconcile differences between the US government technical manuals for reflection
coefficients as a function of angle of incidence and to evaluate the corresponding
reflected scaled impulses, and
5. to provide a technical basis for selecting values of erosion strain suitable for the blast
analysis of reinforced concrete components.

1.4 Notation and Definitions


A number of terms are used throughout this report and are defined here:
Notation
b
bw
Cr

c
c fluid
cp

csolid
cv
D
de
d ( c )
d ( t )

E
e
Fc
Ff

fc

fl
fn
G
Gf
G vpf
h

Ir
Is
J1
J2
J3
K

Definition
Wave form parameter of the Friedlander equation
Width of the stem (web) of a concrete
Reflection coefficient
Speed of sound
Speed of sound in fluids
Specific heats at constant pressure
Speed of sound in solids
Specific heats at constant volume
Detonation velocity
Effective depth from the top of a reinforced concrete beam to the centroid of
the compressive steel
Damage parameter for compression; CSCM model
Damage parameter for tension; CSCM model
Youngs modulus
Specific internal energy
Hardening cap function; CSCM model
Shear failure surface function; CSCM model
Compressive strength of concrete
Effective confining pressure
Natural cyclic frequency
Shear modulus
Fracture energy
Viscoplastic fracture energy
Enthalpy
Reflected impulse
Incident impulse
First invariant of the stress tensor
Second invariant of the deviatoric stress tensor
Third invariant of the deviatoric stress tensor
Bulk modulus

Notation
Kf
K
M
m
Pr
Ps
pr
ps
p
p0

pCJ
pt
qs
R

r
r

rd
r0
rs
Sij

s
Tm
Tn
Tr
t
ta
to
U
u
up

V
Vn

v
vs

Definition
Material constant for the Tuler-Butcher failure criterion
Stiffness
Mach number
Mass
Reflected peak overpressure
Incident peak overpressure
Reflected overpressure
Incident overpressure
Pressure
Ambient pressure
CJ detonation pressure
Hydrostatic tensile pressure
Dynamic pressure
Standoff distance
Radial distance from the center of detonation
Radial expansion of a shock front normalized by charge radius
Radial expansion of a front of detonation products normalized by charge radius
Damage threshold; CSCM model
Initial damage threshold; CSCM model
Deviatoric stress tensor
Entropy
Melting temperature
Period
Room temperature
Time
Arrival time
Positive phase duration
Shock front velocity
Velocity
Particle velocity
Relative volume
Nominal shear strength of concrete
Particle velocity
Particle velocity for incident shock
6

Notation
W
w
wn
X0
Y
Z

D
N

1
3
eff
effdev
effp
f
ij

ijft
ij
ijp
p

incr
inst
jj
p
p
p
vp

Definition
Weight of an explosive
Adiabatic constant; JWL EOS
Natural circular frequency
Initial location of a cap; CSCM model
Common (base 10) logarithm of a blast parameter
Scaled distance; R W 1/3
Angle of incidence
Angle of incidence for the detachment criteria
Angle of incidence for the von Neumann criteria
Angle of reflection
Strain
Strain rate
Reference strain rate
First principal strain
Third principal strain
Effective strain
Effective deviatoric strain
Effective plastic strain
Accumulated plastic strain to failure
Components of the strain tensor
Components of the failure strain tensor
Components of the strain-rate tensor
Components of the plastic strain-rate tensor
Equivalent plastic strain
Incremental geometric strain
Instantaneous geometric strain
Volumetric strain
Equivalent plastic strain
Equivalent plastic strain rate
Increment of accumulated effective plastic strain
Plastic volume strain
Specific heat ratio
Gruneisen coefficient
Strain-rate effect parameter; CSCM model
7

Notation

0
s

0
1
3
eff
effdev
ijft
ijvp

m
y

c
t

Definition
Cap hardening parameter; CSCM model
Wave length
Compression ratio
Poissons ratio
Density
Initial density
Density for incident shock
Stress
Specified threshold stress for the Tuler-Butcher failure criterian
First principal stress
Third principal stress
Effective stress
Effective deviatoric stress
Components of the failure stress tensor
Components of the viscoplastic stress tensor
Hydrostatic stress
Yield stress
Ductile damage; CSCM model
Brittle damage; CSCM model
Rubin three-invariant reduction factor; CSCM model

1.5 Report Organization


This report consists of eight chapters and six appendices. Chapter 2 verifies and validates a CFD
code for analysis of air blast effects, discusses the modeling of close-in (near field) detonations,
and develops an understanding of the characteristics of incident and reflected overpressures in
the near field.
Chapter 3 summarizes past studies on air-blast parameters, including those that underpin US
government documents, presents results of CFD analysis for air-blast parameters across a very
wide range of scaled distance, and compares CFD results to the predictions of Kingery and
Bulmash (1984).

Chapter 4 mines the CFD data of Chapter 3 and presents recommendations for a family of
polynomials to be used for air-blast calculations across the range of scaled distance 0.0553 Z
40 m/kg1/3 (0.139 Z 100 ft/lb1/3).
Chapter 5 examines incident and reflected overpressures and impulses from the detonation of a
high explosive as a function of angle of incidence. A chart and a table to enable the calculation
of reflection coefficients and reflected scaled impulses as a function of the angle of incidence are
presented.
Chapter 6 discusses erosion strain for blast analysis of concrete components for the purpose of
providing a technical basis for the choice of a value of erosion strain. Single element simulations
are performed to develop recommendations. Blast analysis of a sample reinforced concrete
column is performed to identify the importance of correctly assigning a value of erosion strain
for analysis. A step-by-step procedure is provided to compute a value of concrete erosion strain
for blast analysis of concrete components.
Chapter 7 summarizes the technical contributions of this report and presents key conclusions. A
list of references is presented in Chapter 8.
Six appendices supplement this report. Appendix A introduces the speed of sound and its
calculation in solids and fluids. Appendix B describes the relationships between the ChapmanJouguet (CJ) condition, Hugoniot equations, Rayleigh line and the von Neumann spike.
Appendix C assesses incident and reflected overpressures very close to the face of a charge.
Appendix D presents the Kingery and Bulmash charts used in Chapters 3 and 4. Appendix E
presents values of parameters for the constitutive model of concrete used in Chapter 5. Appendix
F describes the single-degree-of-freedom (SDOF) calculations for strain rate to validate the LSDYNA single element simulations in Chapter 6.

SECTION 2
MODELING NEAR-FIELD DETONATIONS OF HIGH EXPLOSIVES
2.1 Introduction
The need to accurately quantify blast pressure loadings in the near field is important because
near-field detonations represent a common threat for the security design of buildings, bridges and
critical infrastructure. Incident and reflected overpressures in the near-field region are too high to
be measured directly by available pressure transducers, or indirectly using air density, because of
the presence of the fireball, requiring a verified (and validated if possible) computational fluid
dynamics (CFD) code to be used to predict the near-field overpressures.
This chapter examines numerical modeling techniques for detonations of spherical high
explosives and characterizes their effects in the near field. The near-field region is defined here
as the region within which the shock wave is affected by local phenomena (e.g., expansion of
detonation products, afterburning) that are insignificant in the far field. These studies are
intended to provide a) confidence in the numerical modeling tools, and b) estimates of incident
and reflected overpressure histories.
Studies are performed using the CFD codes AUTODYN (ANSYS 2009) and Air3D (Rose 2006).
The early expansion of the detonation products is simulated using radial symmetry in one
dimension (1D) before the model is re-mapped into two (2D) and three dimensions (3D)
immediately prior to a boundary being encountered. The Jones-Wilkins-Lee (JWL) Equation of
State (e.g., Fickett and Davis 1979, Needham 2010) is used to model the expanding explosive;
air is modeled as an ideal gas. The JWL Equation of State (EOS) is described in Section 2.2.
AUTODYN is verified for 1D near-field calculations, and 1D and 2D far-field calculations, in
Sections 2.3.2 and 2.3.3, respectively. AUTODYN is validated for 2D near-field calculations and
2D far-field calculations in Sections 2.3.4 and 2.3.5, respectively. Results of one-dimensional
(1D) propagations of blast waves are presented in Section 2.4 to characterize incident
overpressures in the near field. The influences of mesh size, expansion of detonation products

11

and afterburning on pressures and impulses are investigated. AUTODYN and Air3D predictions
are also contrasted.
Two- and three-dimensional simulations are performed in Sections 2.5 and 2.6, respectively, to
study the influence of alternate boundary conditions (or modeling assumptions), to characterize
incident and reflected overpressures, and to confirm the efficacy of reflecting and transmitting
boundaries in AUTODYN. Results of 1D, 2D and 3D simulations are compared to ensure
consistency. The complexity of the wave field in the Mach stem region is identified. The
applicability of the traditional Friedlander waveform (e.g., Baker 1973, Smith and Hetherington
1994, Cormie et al. 2009), namely, an instantaneous rise to peak overpressure followed by an
exponential decay to the ambient condition, to describe near-field pressure loadings is discussed.
Three-dimensional air blast analysis of sample square and circular rigid columns is performed in
Section 2.7 to evaluate reflected overpressures on the surfaces of the columns and the effects of
column shape on the reflected overpressures.

2.2 Modeling Detonations in AUTODYN


AUTODYN is a general-purpose hydrocode used for a wide variety of applications. It uses finite
volume and finite element methods to solve the governing conservation equations. Outputs are
remapped from 1D to 2D/3D and 2D to 3D. The remapping feature allows use of higher
resolution grid in the analysis of the initial stages of the shock wave expansion (ANSYS 2005).
Remapping in AUTODYN is manual and requires user intervention.
Detonation modeling in AUTODYN is two-step process. The first step involves the early time
expansion of the explosive products in 1D using radial symmetry, which continues until a
reflecting surface is reached. The output of the 1D analysis is then transmitted to the 2D/3D
domain that is generated separately. The analysis is the run until a termination time. Although
modeling detonations in AUTODYN is similar to Air3D, the AUTODYN and Air3D solution
methodologies and capabilities differ. Air3D addresses air-blast only.
Analysis of detonations requires the choice of equations of state. For the analyses described in
this report, TNT is modeled using the JWL EOS and air is modeled as an ideal gas. Although
12

AUTODYN has its own material library with default values for the various EOS parameters, the
values of the JWL EOS parameters listed in column 2 of Table 2-1 are used. These values are
identical to those presented by Dobratz and Crawford (1985). The multi-material Euler-Godunov
solver (Godunov 1959) is used for the 1D calculations. (The Euler-FCT solver cannot be used for
radial (1D) calculations.)
Table 2-1 JWL EOS parameters for TNT (Dobratz and Crawford 1985)
Parameter

Value

Density (kg/m3)

1630

A (GPa)

371.2

B (GPa)

3.231

R1

4.15

R2

0.95

Adiabatic constant, w

0.30

Detonation velocity, D (m/s)

6930

Energy per unit volume, E0 (GPa)

CJ pressure, PCJ (GPa)

21

The JWL EOS is a pressure-volume relationship:

w R1V
w R2V w
e
p A 1
e B 1
e
V
1
RV
R2V

(2-1)

where p is the pressure, V is the relative volume, e is the specific internal energy and A, B, R1 ,
R2 , and w are constants obtained by calibration of test data.
Air is modeled as an ideal gas with an initial pressure of 101.3 kPa by specifying an initial
energy of 0.204106 J/kg. The initial temperature, density and adiabatic constant (specific heat
ratio) were set to 288 K (15oC), 1.225 kg/m3, and 1.4, respectively. The multi-material Eulerian
solver is used for both the explosive and air.

13

One-dimensional analysis is performed until the blast wave reaches the end of the 1D domain.
The analysis is typically interrupted when the explosive has expanded to approximately 10 times
its original volume. At this point in the expansion, the value of the compression ratio, ,
presented in Equation 2-2 is approximately -0.99. AUTODYN uses the compression ratio to
calculate the density of the detonation products; see Equation 2-3. The density of the detonation
products becomes very small if falls below -0.99 and this can lead to numerical difficulties.

1
0

(2-2)

0 (1 )

(2-3)

where and 0 are the current density and initial density, respectively. At large volumetric
ratios, the first two terms on the right side of the JWL EOS (Equation 2-1) become negligible
and the EOS collapses to that of an ideal gas EOS, namely:
p ( 1)

e
0

(2-4)

where all terms have been previously defined. Equating Equation 2-4 with Equation 2-1, =
1+w = 1+0.3 = 1.3 for large volumetric ratios. When the compression ratio reaches a value of
-0.99, the EOS for the expanding detonation products is replaced by the EOS for an ideal gas
with = 1.30 and 0 = 1 10-4 gm/cm3 (ANSYS 2009) and the compression ratio is
correspondingly increased to avoid possible numerical instabilities. No change is made to the
EOS for the gas (air) beyond the detonation products. Defining the reference density of a
material associated with the EOS transition is described below based on a personal
communication (ANSYS 2013).
1. When users fill a material to a part in AUTODYN, they will be always asked
about the initial conditions: density, specific internal energy and initial
velocity. The initial material density defaults to the reference density. The
specific internal energy is calculated automatically from the internal energy
per volume. The reference density is equal to the material density in most
cases.
14

2. When TNT detonates, its specific internal energy decreases and the element
pressure increases. Internal energy per volume is not used in the EOS
calculations and thus the reference density does not play a key role in the EOS
calculations.
3. When the reference density is changed by converting the JWL EOS to the
ideal gas EOS, AUTODYN will not change the material density, specific
internal energy, pressure, and other physical variables. The program only
changes the compression ratio, .
Figure 2-1 presents the JWL EOS for TNT using linear and log scales. The legend in each panel
of the figure identifies the terms on the right hand side of Equation 2-1: A, JWL is the first
term, B, JWL is the second term and Gas, JWL is the third term. The line Total is the sum
of these three terms. The contributions of the first two terms, A and B, are significant for relative
volume of 3 and smaller, and insignificant for relative volumes of 5 and greater, where the
relative volume is the ratio of the volume of the expanding explosive to its original volume.

a. Pressure versus relative volume, linear scale

Figure 2-1 JWL and ideal gas EOS for TNT


15

b. Pressure versus relative volume, log (base 10) scale

Figure 2-1 JWL and ideal gas EOS for TNT (cont.)

2.3 Verification and Validation of AUTODYN


2.3.1 Introduction
Verification and validation of numerical codes is standard practice in computational solid
mechanics. Verification is the process of determining that a computational model accurately
represents the underlying mathematical model and its solution. Validation is the process of
determining the degree to which a model is an accurate representation of the real world, as
determined by physical experiments, from the perspective of the intended uses of the model.
Numerical codes, such as the computational fluid dynamics code AUTODYN, should not be
used unless they have been verified and validated. ASME (2006) presents guidance on
verification and validation for computational solid mechanics and that guidance is applied to
AUTODYN.
In this section, AUTODYN is 1) verified using independent numerical predictions and analytical
calculations reported by Needham (2010) for 1D analysis in the near and intermediate fields

16

(Section 2.3.2), 2) verified by cross-code comparisons for 1D and 2D analysis in the far field
(Section 2.3.3), 3) validated in 2D using measurements of normally reflected scaled impulse
reported by Goodman (1960) and Huffington and Ewing (1985) in the near field (Section 2.3.4),
and 4) validated in 2D using measurements of reflected peak overpressures and impulses
reported by Frost et al. (2008) in the far field (Section 2.3.5). Data are not available in the open
literature for either verification in 2D or validation in the near field in 1D. Calculations of
overpressure history cannot be validated in the near field at this time1 and the 2D near-field
validation uses measurements from impulse plugs (e.g., Huffington and Ewing 1985).
The 1D verification study for near-field detonations and the 1D and 2D verification study for farfield detonations are performed using a spherical charge of TNT. The 2D validation study for
near-field detonations is performed using a spherical charge of Pentolite because the only nearfield test data that could be found are measurements of normally reflected scaled impulse
reported by Goodman and Huffington and Ewing for detonations of this explosive. There are no
(reliable) test data (in the open literature) on reflected peak overpressure in the near field. There
are no near field data for detonations of TNT. The 2D validation study for far-field detonations is
performed using measurements of reflected peak overpressure and impulse reported by Frost et
al. for detonations of spherical charges of C4.
2.3.2 Verification of AUTODYN in 1D for Near-Field Detonations
Needham (2010) reports the results of an analysis of a charge of 18000 kg of TNT, with a radius
of 140 cm. The density of the packed TNT was 1570 kg/m3. The analysis was performed using a
Lagrangian finite difference code, detonation products were modeled using the LSZK Equation
of State (EOS) and air was modeled using the Doan Nickel EOS (Doan and Nickel 1963). The
1

Incident and reflected overpressures cannot be measured in the near field (including inside the fireball) using
commercially available pressure transducers because the temperatures are too high. Transducers capable of
measuring very high pressures operate at temperatures less than approximately 600 K: a) the OMEGA PX1009
sensor can measure pressure up to 34 MPa but at a maximum temperature of 616 K (www.omega.com), b) the PCB
PIEZOTRONICS 112A05 sensor can measure pressure up to 34 MPa but at a maximum temperature of 589 K
(www.pcb.com), and c) the PAINE 211-55-010 sensor can measure pressure up to 206 MPa but at a maximum
temperature of 589 K (www.paineelectronics.com). CFD analysis of an explosion in free air predicts that the fireball
expands to a scaled distance of approximately 0.8 m/kg1/3 (2 ft/lb1/3), or a shock front radial expansion of
approximately 15, or an incident overpressure of approximately 2 MPa. The predicted temperature at a scaled
distance of 0.8 m/kg1/3 is greater than 1000 K, with much higher temperatures predicted at smaller values of scaled
distance.

17

results presented in Needham were digitized to enable a direct comparison with results computed
using AUTODYN. Needham reported results using centimeters for distance. This unit is adopted
in this chapter where comparisons of Needham and CFD data are made.
A 1D AUTODYN model of the charge described above is prepared using the JWL EOS for the
detonation products and the ideal gas EOS for air. The values of the parameters used for the JWL
EOS, presented in Table 2-1, are those reported in the LLNL Explosives Handbook (Dobratz and
Crawford 1985). (Needham did not report the values he used for the EOS.) The LLNL density
for packed TNT is 1630 kg/m3, and that is 4% greater than that assumed by Needham. To use a
consistent set of properties for the JWL EOS, and a charge shape identical to that of Needham to
enable reporting at the same values of radial expansion provided in Needham, the TNT weight
for the AUTODYN analysis is 18735 kg. The 4% increase in weight will have a minor effect on
the overpressure histories but small differences between AUTODYN predictions and results
reported by Needham are inevitable.
Needham (2010) reported hydrodynamic parameters at different values of shock front radial
expansion, r : 1.1, 2.4, 2.6, 4.5, 11.7, 26 and 34. Needham and AUTODYN results are presented
in Figures 2-2, 2-3, 2-4, 2-5, 2-7, 2-9 and 2-10. Additional data are presented in Figures 2-6 and
2-8 for shock front radial expansions of 8 and 18 because the hydrodynamic parameters vary
greatly between shock front radial expansions of 4.5 and 11.7, and 11.7 and 26, respectively. The
hydrodynamic parameters reported in each figure are relative overpressure ( p / p0 1 ), relative
over density ( / 0 1 ), and particle velocity, v, in units of km/second. Relative overpressure is
presented as the total pressure, p, normalized by the ambient air pressure ( p0 =101.3 kPa) minus
1, and relative over density is presented as the total density normalized by the ambient
atmospheric density ( 0 = 1.225 kg/m3) minus 1. The AUTODYN calculations were performed
using an Eulerian solver and a mesh with a cell dimension of 20 mm.

18

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-2 Hydrodynamic parameters at r = 1.1 for a 20 mm mesh
19

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-3 Hydrodynamic parameters at r = 2.4 for a 20 mm mesh
20

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-4 Hydrodynamic parameters at r = 2.6 for a 20 mm mesh
21

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-5 Hydrodynamic parameters at r = 4.5 for a 20 mm mesh
22

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-6 Hydrodynamic parameters at r = 8 for a 20 mm mesh
23

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-7 Hydrodynamic parameters at r = 11.7 for a 20 mm mesh
24

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-8 Hydrodynamic parameters at r = 18 for a 20 mm mesh
25

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-9 Hydrodynamic parameters at r = 26 for a 20 mm mesh
26

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-10 Hydrodynamic parameters at r = 34 for a 20 mm mesh
27

Needham reported that the detonation velocity, composed of a sound speed (see Appendix A) at
the detonation front of 5.2 km/sec and a peak material velocity of 1.8 km/sec, was 7 km/sec,
immediately prior to completion of the detonation (i.e., shock front radial expansion, r , of 1.0).
Figure 5.1 of Needham shows that the material velocity at this time instant decays from 1.8
km/sec at the shock front to zero at a radius of 65 cm and that the density (1300 kg/m3) and
pressure (4.7 MPa) are constant for a radial distance of less than 65 cm where the particles are
stationary.
At a shock front radial expansion of 1.1, the normalized pressures and densities are most similar
but the AUTODYN peak velocity is substantially smaller than that of Needham of 7.4 km/sec,
which is attributed to the coarseness of the chosen mesh (see mesh convergence study later in
this section). The peak material velocity of 7.4 km/sec, which is recovered with a much finer
mesh (see mesh convergence study later in this section), has increased from 1.8 km/sec at a
shock front radial expansion of 1.0. The shock front velocity, which is the sum of the sound
speed (0.34 km/s) in air and the peak material velocity, is 7.7 km/s and greater than the
detonation velocity. In panel a) of Figure 2-2, it can be seen that the head of the rarefaction wave
is at approximately r = 130 cm and 10 cm inside the radius of the charge prior to detonation.
At a shock front radial expansion of 2.4, the AUTODYN and Needham results are similar, with
the greatest differences observed for relative overpressure in which it is seen that the rarefaction
wave propagates fastest in AUTODYN. In panel a) of Figure 2-3, it is evident that the head of
the rarefaction wave has reached the point of detonation (r = 0 cm) in the AUTODYN analysis
but only r = 35 cm in the Needham analysis.
At a shock front radial expansion of 2.6, the hydrodynamic parameters for pressure and density
are similar for distances of r 80 cm and at all distances for particle velocity.
At shock front radial expansions of 11.7, 26 and 34, the differences between the AUTODYN and
Needham results are by-and-large small. A difference is observed in the relative overdensity at a
distance of 1400 cm at a shock front radial expansion of 11.7 (i.e., shock front at 1638 cm from
the center of the charge), where a 20 mm mesh cannot resolve the sharp density spike.

28

Figure 2-11 is provided to observe the relationship between the pressure, density and velocity for
shock front radial expansions of 4.5, 8, 11.7, 18 and 26. For a shock front radial expansion of
4.5, the local increase at approximately r = 550 cm is due to the front of the detonation product,
at which the pressure history is discontinuous. The peak particle velocity is at r 500 cm,
immediately behind the front of the detonation products. The tail of the rarefaction wave is at the
point of the peak particle velocity. Similar observations are made for all other shock front radial
expansions. For a shock front radial expansion of 26, the velocity at the front of the detonation
products (located by an open circle) is nearly zero, which is associated with the discontinuities at
r 2200 cm in the pressure and density profiles, respectively, also shown with open circles.

a. r = 4.5

b. r = 8
Figure 2-11 Hydrodynamic parameters for a 20 mm mesh in AUTODYN
29

c. r = 11.7

d. r = 18

e. r = 26
Figure 2-11 Hydrodynamic parameters for a 20 mm mesh in AUTODYN (cont.)
30

Needham estimated the acceleration of the air immediately beyond the face of the charge at the
time the detonation is complete using
dv
1 dp

0 dr
dt

(2-5)

where v is the particle velocity, 0 is the ambient density of air, p is pressure and r is radius. For
p equal to the average of the detonation pressure2 (approximately 21000 MPa) and ambient air
pressure, an ambient air density of 1.225 kg/m3, and a value of dr equal to 0.01 m (1 cm), the
particle acceleration is approximately 8.6 1011 m/sec2, or 8.6 1010 g. Figure 2-12 plots
acceleration histories calculated using a cell size of 1 mm at five distances from the center of the
explosive; a radius of 140 cm represents the face of the charge. The calculated peak accelerations
are smaller than those computed using Equation 2-5, in part due to cell size, as noted above.

Figure 2-12 Acceleration histories at discrete points along the 1D expansion


Needham also showed the process of the formation of a secondary shock wave. The (expanding)
rarefaction waves moving toward the center of the charge reduce the pressures and densities
behind the front of the expanding detonation products. The particle velocity behind the front of
the expanding detonation products also decreases and becomes negative. At this time, particles
moving at negative velocities lead to shock waves propagating toward the center of the charge.
These shock waves are shown in Figure 2-9, where the front of the shock waves is at about r
2

The calculation of the detonation pressure is discussed in Appendix B.

31

1000 cm and 1200 cm in the AUTODYN and Needham analyses, respectively. The shock wave
moving inward is then reflected outward from the center of the charge, which is typically
described as the secondary shock wave. The formation of this secondary shock wave is seen in
Figure 2-10 for the shock front radial expansion of 34. The secondary shock front is at r 650
cm and 900 cm in the AUTODYN and Needham analyses, respectively.
In summary, the Needham and AUTODYN results are close but not identical. This is attributed
to differences in meshing, the choice of solvers and equations of state, and initial choice of TNT
density. The fact that the relative overpressures, overdensities, and particle velocities are similar
provides confidence in the use of AUTODYN. The importance of choice of cell size, especially
for small values of shock front radial expansion, is discussed below in this section and Sections
2.3.3, 2.4, 2.5 and 2.6.
Figures 2-13 through 2-21 present histories of the hydrodynamic parameters at different values
of shock front radial expansion. The normalized density history in Figure 2-14 shows three peaks
at 0.59 msec, 0.65 msec, and 1.5 msec, which are associated with the arrival of the shock front,
arrival of the front of the detonation products, and the peak outflow of detonation products
(affected by rarefaction waves moving toward the center of the charge), respectively. The peak in
the normalized pressure history in Figure 2-14 (and Figures 2-13, 2-15, 2-16, 2-17, 2-18, 2-19, 220, 2-21) is associated with the arrival of the shock front and not the expanding detonation
products.

32

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-13 Histories of hydrodynamic parameters at r = 1.1 for a 20 mm mesh

33

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-14 Histories of hydrodynamic parameters at r = 2.4 for a 20 mm mesh

34

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-15 Histories of hydrodynamic parameters at r = 2.6 for a 20 mm mesh

35

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-16 Histories for hydrodynamic parameters at r = 4.5 for a 20 mm mesh

36

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-17 Histories of hydrodynamic parameters at r = 8 for a 20 mm mesh

37

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-18 Histories of hydrodynamic parameters at r = 11.7 for a 20 mm mesh

38

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-19 Histories of hydrodynamic parameters at r = 18 for a 20 mm mesh

39

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-20 Histories of hydrodynamic parameters at r = 26 for a 20 mm mesh

40

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-21 Histories of hydrodynamic parameters at r = 34 for a 20 mm mesh

41

A mesh (cell size) convergence study is performed for small values of shock front radial
expansion: 0.5, 0.9, 1.1, 2.4 and 2.6. Results are presented in Figures 2-22 through 2-26,
respectively. For values less than 1, the shock front lies within the explosive. Data from
Needham (2010) are presented here for shock front radial expansions of 1.1, 2.4 and 2.63. The
cell (mesh) sizes chosen for AUTODYN analysis were 20 mm (see Figures 2-2 through 2-10), 10
mm, 5 mm and 1 mm. A comparison of results for the radial expansions greater than 1 shows
that the finer the AUTODYN mesh, the closer the results are to those reported by Needham. The
influence of cell size is pronounced for particle velocity at a shock front radial expansion of 1.1.
The speed of propagation of the rarefaction wave is better resolved with the finer meshes as seen
in Figure 2-25a. For the mesh with 20 mm cells, AUTODYN predicts the rarefaction wave has
reached the point of detonation at a shock front radial expansion of 2.4. For the 1 mm cell size,
AUTODYN predicts that the rarefaction wave has reached a distance, r, equal to 25 cm from the
point of detonation at a shock front radial expansion of 2.4, which correlates better with the
Needham prediction of r = 35 cm.
The importance of cell size for the correct modeling of the propagation of the shock front
through the explosive is seen clearly in Figures 2-22 and 2-23. The rapid changes in pressure,
density and particle velocity across the shock front are not captured with the meshes constructed
with 10 mm and 20 mm cells.
In Figures 2-22 and 2-23, the detonation fronts are identified at r 68 cm and 125 cm,
respectively. The peak relative overpressure at the detonation front is approximately 180000 for
the 1 mm mesh. The relative overdensity of the charge near its center after detonation is 1090,
which is 18% less than that of a cold (unreacted) TNT ( 1330 (= 1630 kg/m3/1.225 kg/m3 1,
where 1630 and 1.225 are the TNT packing density and ambient air density, respectively). At the
detonation front, the relative overdensity is approximately 1650, which is 24% greater than that
of cold TNT, which shows that the detonation products are not very compressible because the
3

Numerical studies to investigate incident peak overpressure and reflected peak overpressure in very
close proximity to the face of a charge, 0.981 r 1.034, are presented in Appendix C. Those studies
are not integrated into this presentation because Needham did not report results in this region. A charge
with a radius of 150 mm is used in Appendix C and results are reported at distances of 147 r 155 mm,
namely, 0.981 r 1.034 and 0.517 Z 0.545 m/kg1/3.

42

pressure at the detonation front ( 180 kbars) is extremely high. The peak velocity is
approximately 1.6 km/sec for the finest mesh of 1 mm in panel c) of Figure 2-23, which is
similar to that in the Needham analysis (=1.8 km/sec).
A mesh convergence study for overpressure and impulse is also performed for the 18735 kg
charge used in this section at shock front radial expansions of 1.025, 1.1, 2.4, 4.5 and 11.7; Z =
0.054, 0.058, 0.13, 0.24 and 0.62 m/kg1/3, respectively, where Z is the scaled distance, defined in
Section 1.4. The radial expansion of r = 1.025 is chosen in preference to r = 1.0 because the
particle acceleration and velocity very close to the face of the charge change rapidly but both
stabilize at a radial expansion r = 1.025 (see Section 3.5). Cell sizes of 400, 200, 100, 40, 20, 10,
5 and 2 mm are considered. Results are presented in Figures 2-27 and 2-28. For the five shock
front radial expansions of 1.025, 1.1, 2.4, 4.5 and 11.7, the overpressure histories converge for
cell sizes of 5, 5, 10, 20 and 200 mm, respectively, (r/287, r/308, r/336, r/315 and r/82), where
convergence is defined as a result within 20% of the true result calculated with the smallest
cell size (2 mm here). For impulse, the results converge for cell sizes of 10, 10, 10, 40 and 400
mm, respectively (r/144, r/154, r/336, r/158 and r/41). The overpressure histories, especially for
Figures 2-27c and 2-27d are quite different from the Friedlander curve (described in Section 2.1)
due to the influence of expanding detonation products, which are discussed in more detail in
Sections 2.4 and 2.5.
We thus consider AUTODYN to be verified for 1D calculations of overpressure and impulse for
near-field detonations.

43

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-22 Mesh convergence study at r = 0.5
44

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-23 Mesh congervence study at r = 0.9
45

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-24 Mesh convergence study for r = 1.1
46

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-25 Mesh convergence study for r = 2.4
47

a. Pressure ( p0 = 0.1 MPa)

b. Density ( 0 = 0.001225 g/cm3)

c. Particle velocity, v
Figure 2-26 Mesh convergence study for r = 2.6
48

a. r = 1435 m (35 mm from face of charge); Z = 0.054 m/kg1/3; r =1.025

b. r = 1540 mm (140 mm from face of charge); Z = 0.058 m/kg1/3; r =1.10


Figure 2-27 Overpressure histories at distance, r, from the center of a spherical charge of
18735 kg of TNT; JWL EOS

49

c. r = 3360 mm (1960 mm from face of charge); Z = 0.013 m/kg1/3; r =2.4

d. r = 6300 mm (4900 mm from face of charge); Z = 0.24 m/kg1/3; r =4.5


Figure 2-27 Overpressure histories at distance, r, from the center of a spherical charge of
18735 kg of TNT; JWL EOS (cont.)

50

e. r = 16380 mm (14980 mm from face of charge); Z = 0.62 m/kg1/3; r =11.7


Figure 2-27 Overpressure histories at distance, r, from the center of a spherical charge of
18735 kg of TNT; JWL EOS (cont.)

51

a. r = 1435 m (35 mm from face of charge); Z = 0.054 m/kg1/3; r =1.025

b. r = 1540 mm (140 mm from face of charge); Z = 0.058 m/kg1/3; r =1.10


Figure 2-28 Impulse histories at distance, r, from the center of a spherical charge of 18735
kg of TNT; JWL EOS

52

c. r = 3360 mm (1960 mm from face of charge); Z = 0.013 m/kg1/3; r =2.4

d. r = 6300 mm (4900 mm from face of charge); Z = 0.24 m/kg1/3; r =4.5


Figure 2-28 Impulse histories at distance, r, from the center of a spherical charge of 18735
kg of TNT; JWL EOS (cont.)

53

e. r = 16380 mm (14980 mm from face of charge); Z = 0.62 m/kg1/3; r =11.7


Figure 2-28 Impulse histories at distance, r, from the center of a spherical charge of 18735
kg of TNT; JWL EOS (cont.)
2.3.3 Verification of AUTODYN in 1D and 2D for Far-Field Detonations
Browning et al. (2013) simulated an 8000 lb (3629 kg) detonation of TNT in 2D using LSDYNA (LSTC 2013) and CTH (McGlaun et al. 1990). Three cell sizes were used for the
simulations: 1, 2 and 3 in. (25.4, 50.8 and 76.2 mm, respectively). They reported results at scaled
distances of 3, 4 and 5 ft/lb1/3 (1.190, 1.587 and 1.987 m/kg1/3, respectively) as noted in Table 22. The Browning et al. calculations of incident peak overpressure, incident impulse, and arrival
time were repeated using AUTODYN. Table 2-3 presents the values of the JWL parameters for
modeling TNT used by Browning et al. and in this verification study. (The values are essentially
identical to those provided in Table 2-1.) Values for modeling air are presented in Table 2-4.

54

Table 2-2 Monitoring locations used by Browning et al. (2013)


Range, ft [m]

Scaled distance, ft/lb1/3 [m/kg1/3]

60 [18.3]

3.00 [1.190]

80 [24.4]

4.00 [1.587]

100 [30.5]

5.00 [1.984]

Table 2-3 JWL EOS parameters for modeling TNT used by Browning et al. (2013)
Parameter

Value

Charge radius (in) [mm]

31.89 [810]

Density (lb/in3) [kg/m1/3]

0.058887 [1630]

A (psi) [GPa]

5.384107 [371.2]

B (psi) [GPa]

4.685105 [3.230]

R1

4.15

R2

0.95

Adiabatic constant, w

0.3

Detonation velocity, D (in/s) [m/s]

2.728105 [6929]

Energy per unit volume, E0 (psi) [GPa]

1.015106 [6.998]
3.046106 [21]

CJ pressure, PCJ (psi) [GPa]

Table 2-4 Values of input parameters for modeling air for far-field verification
Parameter

This study

Browning et al. (2013)

EOS

Ideal gas

Linear polynomial (ideal gas)

Density (lb/in3) [kg/m1/3]

4.42610-5 [1.225]

4.42010-5 [1.223]

Specific heat ratio,

1.4

1.4

Energy per unit volume, E0 (psi) [kPa]

36.738 [253.3]

36.740 [253.3]

Figure 2-29 presents AUTODYN histories of incident overpressure and impulse in 1D at scaled
distances of 3, 4 and 5 ft/lb1/3 and cell sizes of 1, 2 and 3 in. The incident overpressure and
impulse histories are effectively independent of cell size. Figure 2-30 enables a comparison of

55

the AUTODYN predictions of incident peak overpressures, incident impulses and arrival times
for the three cell sizes. The influence of cell size on these results is not significant.
Figures 2-31 and 2-32 summarize numerical results from the 1D AUTODYN, 2D LS-DYNA
(Browning et al.) and 2D CTH (Browning et al.) simulations for cell sizes of 1, 2 and 3 in.
Incident peak overpressures are provided in Tables 2-5, 2-6 and 2-7 for cell sizes of 1, 2 and 3 in,
respectively. Maximum values of incident impulse are provided in Tables 2-8, 2-9 and 2-10. The
LS-DYNA and CTH calculations presented in Tables 2-5 through 2-10 are from Browning et al.
AUTODYN, LS-DYNA and CTH provide comparable values of incident peak overpressure and
impulse in the far field. Small differences are expected, even in the far field and for the same cell
size, because different algorithms and solvers are employed in these codes.
We thus consider AUTODYN to be verified for 1D and 2D calculations of overpressure and
impulse for far-field detonations.
Table 2-5 Results for incident peak overpressure; cell size of 1 in; units of psi
Z (ft/lb1/3)

AUTODYN

LS-DYNA

CTH

94.1

86.2

79.9

46.6

42.8

39.0

27.8

25.8

23.2

Table 2-6 Results for incident peak overpressure; cell size of 2 in; units of psi
Z (ft/lb1/3)

AUTODYN

LS-DYNA

CTH

88.3

82.7

79.3

44.3

41.7

38.0

26.7

25.3

22.6

56

Table 2-7 Results for incident peak overpressure; cell size of 3 in; units of psi
Z (ft/lb1/3)

AUTODYN

LS-DYNA

CTH

83.8

80.0

76.2

42.5

40.7

37.5

25.8

24.9

22.4

Table 2-8 Results for incident impulse; cell size of 1 in; units of psi-ms
Z (ft/lb1/3)

AUTODYN

LS-DYNA

CTH

264

236

215

206

191

175

176

164

151

Table 2-9 Results for incident impulse; cell size of 2 in; units of psi-ms
Z (ft/lb1/3)

AUTODYN

LS-DYNA

CTH

259

236

214

203

191

175

174

164

151

Table 2-10 Results for incident impulse; cell size of 3 in; units of psi-ms
Z (ft/lb1/3)

AUTODYN

LS-DYNA

CTH

255

237

214

201

191

174

172

164

151

57

Incident overpressure

Incident impulse

(a) Z = 3 ft/lb1/3

(b) Z = 4 ft/lb1/3

(c) Z = 5 ft/lb1/3
Figure 2-29 Incident overpressure and impulse histories from AUTODYN; Z = 3, 4 and 5
ft/lb1/3; cell sizes = 1, 2 and 3 in
58

(a) Incident peak overpressure

(b) Incident impulse

(c) Arrival time


Figure 2-30 Incident peak overpressure and impulse from AUTODYN; Z = 3, 4 and 5
ft/lb1/3; cell sizes = 1, 2 and 3 in
59

(a) 1 in cell size

(b) 2 in cell size

(c) 3 in cell size


Figure 2-31 Summary of AUTODYN (this study) and LS-DYNA and CTH (Browning et
al.) results for incident peak overpressure
60

(a) 1 in cell size

(b) 2 in cell size

(c) 3 in cell size


Figure 2-32 Summary of AUTODYN (this study) and LS-DYNA and CTH (Browning et
al.) results for incident impulse
61

2.3.4 Validation of AUTODYN in 2D for Near-Field Detonations


Two-dimensional (2D) numerical analysis is performed to validate AUTODYN for calculations
of reflected overpressure and impulse in the near field. There is a limited body of welldocumented test data with which to validate computational fluid dynamics codes.
This validation exercise uses measured values of normally reflected scaled impulse reported by
Goodman (1960) and Huffington and Ewing (1985) for detonations of spherical charges of 50/50
Pentolite. No peer reviewed data were found in the open literature for TNT. Goodman compiled
and presented experimental data for incident (side-on) and normally reflected overpressure and
impulse for detonations of spherical 50/50 Pentolite in free air. No data were reported for
incident impulse and normally reflected overpressure for Z < 0.4 m/kg1/3. Data for incident peak
overpressure were reported for small values of scaled distance but these were inferred from
optical measurements of shock front velocity and not measured directly. The values of normally
reflected scaled impulse were measured using an impulse plug (e.g., Johnson et al. 1957,
Huffington and Ewing 1985) and are used here to validate AUTODYN. Goodman presented
normally reflected scaled impulse data for Z = 0.0742, 0.0869, 0.127, 0.150, 0.174 and 0.198
m/kg1/3 (0.187, 0.219, 0.320, 0.378, 0.438 and 0.500 ft/lb1/3, respectively), based on 9, 9, 7, 26, 8
and 37 tests, respectively (see Table II of the appendix in Goodman). The values of normally
reflected scaled impulse for Z = 0.08, 0.12, 0.16 and 0.20 m/kg1/3 were interpolated from these
data for comparison with AUTODYN predictions.
Huffington and Ewing report measured normally reflected impulses for spherical charges of
50/50 Pentolite from 57 experiments at scaled distances between 0.06 and 0.20 m/kg1/3 (0.15 and
0.50 ft/lb1/3, respectively). The masses of the charges used for these experiments were 0.5, 1 and
2 lbs (0.227, 0.454 and 0.907 kg, respectively). Huffington and Ewing presented 13, 16, 15 and
10 test results for Z = 0.08, 0.12, 0.16 and 0.20 mkg1/3, respectively (see Table 1 in Huffington
and Ewing), and the mean results at each value of scaled distance are used to validate
AUTODYN.
Pentolite is modeled using the JWL EOS and air is modeled as an ideal gas. Four charge weights
of 0.5, 1, 22.68 and 18735 kg, with a packing density of 1700 m/kg1/3, are analyzed to observe
62

the effect of charge weight; see Table 2-11. The JWL parameters for 50/50 Pentolite used in the
AUTODYN simulations are those reported in Dobratz and Crawford (1985); see Table 2-12.
Table 2-11 Properties for 50/50 Pentolite and C4
Parameter

Pentolite

C4

Weight (kg)

0.5

1.0

22.68

18735

0.5

Radius (mm)

41.25

52.00

147.13

1380.5

53

42

Density (kg/m3)

1700

1700

1700

1700

1601

1601

Table 2-12 JWL EOS parameters for 50/50 Pentolite and C4 (Dobratz and Crawford 1985)
Parameter

Pentolite

C4

1700

1601

A (GPa)

540.94

609.77

B (GPa)

9.3726

12.95

R1

4.5

4.5

R2

1.1

1.4

Adiabatic constant, w

0.35

0.25

Detonation velocity, D (m/s)

7530

8193

Energy per unit volume, E0 (GPa)

8.1

9.0

CJ pressure, PCJ (GPa)

25.5

28

Density (kg/m3)

One-dimensional (1D) radial analysis for the 1kg Pentolite is performed using AUTODYN until
the shock wave reaches a reflecting boundary. Five cell sizes of r/125, r/250, r/500, r/1000 and
r/2000 are used for all four scaled distances, where r is the distance of a monitoring location
from the center of the charge. Note that these five cell sizes produce differences of 4.8%, 1.0%,
1.0%, 0.13% and 0.13%, respectively, from the exact charge weight (= 1 kg), but these will have
a minor effect on the overpressure histories. The 1D analysis results calculated using the r/2000
cell size are mapped into a 2D air domain at (x, y) = (0, 0) for analysis using three cell sizes:
r/125, r/250 and r/500. The dimension of the 2D domain is r r, where the lower horizontal
boundary is axially symmetric, the left vertical and the top horizontal boundaries are transmitting

63

planes, and the right vertical boundary is perfectly reflecting plane. Energy will seldom escape
across the left transmitting (upstream) boundary because the pressure waves initially propagate
radially from the source and away from this boundary. The monitoring location is on the right
perfect-reflecting boundary, at (x, y) = (r, 0).
Figure 2-33 presents simulation results for 1kg of Pentolite at Z = 0.08 m/kg1/3. Figures 2-33a
and 33b show the converged 1D results using a cell size of r/2000. The reflected overpressure
and impulse histories in Figures 2-33c and 2-33d, respectively, are similar for all three (2D) cell
sizes. Identical observations are made for Z = 0.12, 0.16 and 0.20 m/kg1/3, as shown in Figures
2-34, 2-35, and 2-36, respectively. Analyses for Pentolite masses of 0.5, 22.68 and 18735 kg are
also performed using 1D and 2D cell sizes of r/2000 and r/500, respectively.
Figure 2-37 enables a comparison of the AUTODYN-calculated normally reflected scaled
impulses, with the smallest cell sizes for four scaled distances and four charge weights, and the
measurements of Goodman and of Huffington and Ewing. See Table 2-13. For all four scaled
distances, the AUTODYN calculations are in good agreement with the measured data. The effect
of charge weight on scaled impulse is insignificant.
We thus consider AUTODYN to be validated for 2D calculations of reflected overpressure and
impulse for near-field detonations.
Table 2-13 Normally reflected scaled impulse from AUTODYN calculations and
measurements of Goodman (1960) and Huffington and Ewing (1985); units of
MPa-ms/kg1/3
Z (m/kg1/3)

AUTODYN

Goodman

Huffington
and Ewing

0.5 kg

1 kg

22.68 kg

18735 kg

0.08

38.95

38.95

38.93

38.94

41.76

42.25

0.12

18.16

18.15

18.11

18.13

19.49

18.92

0.16

10.53

10.54

10.52

10.54

11.47

10.81

0.2

6.880

6.882

6.873

6.84

7.14

7.21

64

(a) Incident overpressure

(b) Incident impulse

(c) Reflected overpressure

(d) Reflected impulse

Figure 2-33 Incident and normally reflected overpressure and impulse histories calculated
using AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.08 m/kg1/3; (Z = 0.20
ft/lb1/3); r = 1.52

65

(a) Incident overpressure

(b) Incident impulse

(c) Reflected overpressure

(d) Reflected impulse

Figure 2-34 Incident and reflected normally overpressure and impulse histories calculated
using AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.12 m/kg1/3; (Z = 0.30
ft/lb1/3); r = 2.28

66

(a) Incident overpressure

(b) Incident impulse

(c) Reflected overpressure

(d) Reflected impulse

Figure 2-35 Incident and normally reflected overpressure and impulse histories calculated
using AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.16 m/kg1/3; (Z = 0.40
ft/lb1/3); r =3.04

67

(a) Incident overpressure

(b) Incident impulse

(c) Reflected overpressure


(d) Reflected impulse
Figure 2-36 Incident and normally reflected overpressure and impulse histories calculated
using AUTODYN for 50/50 Pentolite of 1 kg; Z = 0.20 m/kg1/3; (Z = 0.50
ft/lb1/3); r = 3.80

Figure 2-37 Comparison between AUTODYN calculations and measurements of Goodman


(1960) and Huffington and Ewing (1985) for normally reflected scaled impulses
68

2.3.5 Validation of AUTODYN in 2D for Far-Field Detonations


Two-dimensional (2D) numerical analysis for far-field detonations is performed to validate the
code for the calculations of reflected overpressures and impulses in the far field. This validation
study uses measurements of normally reflected peak overpressures and impulses reported by
Frost et al. (2008) for spherical charges of C4. AUTODYN Analysis is performed and values of
normally reflected peak overpressures and scaled impulses at scaled distances, Z, of 1.2 and 1.51
m/kg1/3 are reported and compared.
Frost et al. presented measured data of reflected peak overpressures and impulses for normal
incidences and angles of incidence for detonations of spherical charges of 1 kg and 0.5 kg of C4,
1.2 m from the center of the charge (Z = 1.2 and 1.51 m/kg1/3, respectively). The weight, radius
and packing density of both C4 charges are presented in Table 2-11. The results presented by
Frost et al. for normal incidence were digitized to enable a direct comparison with results
computed using AUTODYN.
The JWL parameters for C4, from Dobratz and Crawford (1985), are used for the AUTODYN
simulations, and are presented in Table 2-12. Cell sizes of r/1200 and r/2400 (1 mm and 0.5 mm,
respectively) are used for 1D radial analysis because these enable exact modeling of the charge
radii of 53 mm and 42 mm, for the C4 weights of 1.0 kg and 0.5 kg, respectively, and the
packing density of 1601 kg/m3. The 1D analysis results calculated using the r/2400 cell size are
used for 2D analysis, where four cell sizes of r/125, r/250, r/500 and r/1000 are considered.
Modeling of the 2D domain is similar to that described in the previous section.
Figure 2-38 presents simulation results for Z = 1.2 m/kg1/3 (1 kg C4). The incident overpressure
and impulse histories are virtually identical for the two cell sizes, as shown in Figures 2-38a and
2-38b, respectively. For normally reflected peak overpressure, the use of cell sizes of r/500 and
r/1000 lead to similar results, as shown in Figure 2-38c, where the difference is less than 10%.
Figure 2-38d shows the normally reflected impulse histories, which are essentially identical for
all four cell sizes. Identical observations are made for Z = 1.51 m/kg1/3 (0.5 kg C4), as shown in
Figure 2-39.

69

Figure 2-40 enables a comparison between AUTODYN results calculated using the r/1000 cell
size and Frost et al. predictions. The calculated reflected peak overpressure for 1.51 m/kg1/3 is
somewhat (32%) smaller than that predicted by Frost et al., but the value for 1.2 m/kg1/3 is close
to that predicted by Frost et al. Further, the calculated reflected scaled impulses for both scaled
distances and the Frost et al. predictions are in good agreement.
We thus consider AUTODYN to have been validated for 2D calculations of reflected
overpressure and impulse for far-field detonations.

(a) Incident overpressure

(b) Incident impulse

(c) Reflected overpressure

(d) Reflected impulse

Figure 2-38 Incident and normally reflected overpressure and impulse histories calculated
using AUTODYN for C4 of 1 kg; Z = 1.2 m/kg1/3 (r = 1.2 m)

70

(a) Incident overpressure

(b) Incident impulse

(c) Reflected overpressure


(d) Reflected impulse
Figure 2-39 Incident and normally reflected overpressure and impulse histories calculated
using AUTODYN for C4 of 0.5 kg; Z = 1.51 m/kg1/3 (r = 1.2 m)

Figure 2-40 Comparison between AUTODYN calculations and Frost et al. measurements
(2008) for normally reflected peak overpressure, Pr, and scaled impulse, Ir/W1/3;
Z = 1.2 and 1.51 m/kg1/3
71

2.4 One-Dimensional Blast Wave Propagation


2.4.1 Introduction
A series of 1D studies is performed using AUTODYN and Air3D (Rose 2006). The purpose of
the 1D studies is to characterize the incident overpressure histories in close proximity to the face
of a spherical weapon of 22.68 (50 lb) of TNT. A 22.68 kg weapon is a typical hand-delivered
threat for security design of buildings and infrastructure. The influences of mesh size,
afterburning, and equations of state (EOS) on overpressure history, and temperature in the
vicinity of the face of the charge are studied.
A charge radius of 149 mm corresponding to the 22.68 kg TNT is modeled in 1D and is meshed
with 4, 2, 1, 0.5 and 0.25 mm cells. A TNT packing density of 1636.8 kg/m3 is used to recover
the weight of 22.68 kg with a charge radius of 149 mm. The JWL EOS is used to model the TNT
products and air is modeled as an ideal gas. The multi-material Euler-Godunov solver is used for
the 1D calculations.
2.4.2 Mesh Sensitivity Study
A mesh sensitivity study is performed for simulating 1D wave propagation using the 22.68 kg
charge of TNT. Afterburning is not considered. An h-refinement of the mesh is performed using
five cell sizes: 4 mm, 2 mm, 1 mm, 0.5 mm and 0.25 mm. The choice of cell size influences the
results, over the range considered, for a distance r of less than seven times the charge radius (=
1043 mm), or a radial expansion of 7. Results are presented in Figures 2-41 and 2-42 as a
function of distance, r, from the center of the charge. For distances greater than 416 mm (Z =
0.147 m/kg1/3), results are also compared with data from CONWEP (Hyde 1992) as implemented
in LS-DYNA (LSTC 2013). CONWEP data are available only for Z > 0.147 m/kg1/3. The rate of
decay in overpressure is described by the Friedlander curve (see Equation 1.1) and the waveform
parameter (b in that equation) is back-calculated from the CONWEP values for peak
overpressure, impulse, and positive phase duration. At distances of 152.7 mm, 164 mm, 200 mm
and 250 mm (Z = 0.054, 0.058, 0.071 and 0.088 m/kg1/3, respectively; r = 1.025, 1.1, 1.34 and
1.68, respectively), the overpressures are resolved sufficiently well, as measured by the small
72

differences of generally less than 10% in results for cell sizes of 0.5 mm and 0.25 mm. Consider
Figure 2-41c. The 0.5 mm and 0.25 mm cells capture the spike associated with the arrival of the
shock front (peak overpressure of approximately 40 MPa; see open circle) and the expanding
detonation products (peak overpressure of approximately 77 MPa for the 0.5 mm cell size),
immediately behind the initial peak. Similar observations are made for Figure 2-41d although the
relative amplitudes of the peaks in the overpressure histories have changed because the meshes
are better able to resolve the initial shock front. At a distance of 500 mm (Z = 0.177 m/kg1/3; r =
3.36; Figure 2-41e), the calculated overpressure histories are quite different from CONWEP due
to the effect of the expanding detonation products. At a distance of 1000 mm (Z = 0.353 m/kg1/3;

r = 6.71; Figure 2-41f), the five meshes generated similar results. The distance of 1000 mm
(Figure 2-41f) represents a volume expansion of 302. The shock front dominates the
overpressure history at this radial expansion. The differences between the calculated results and
CONWEP predictions at t = 0.3 to 0.4 ms are also due to the effect of the expanding detonation
products. At distances of 152.7, 164, 200, 250, 500 and 1000 mm ( r = 1.025, 1.1, 1.34, 1.68,
3.36 and 6.71, respectively), the overpressure histories converge for cell sizes of 0.5, 0.5, 0.5, 0.5,
1.0 and 2.0 mm, respectively (or r/306, r/328, r/400, r/500, r/500 and r/500). For the total
impulse, the results converge at larger cell sizes, namely, r/153, r/164, r/100, r/125, r/250 and
r/250, respectively. None of the pressure time series presented in Figures 2-41c through 2-41f are
similar in shape to the widely adopted Friedlander equation. The time series in Figures 2-41a and
2-41b are different because at these small radial expansions (1.025 and 1.10, respectively), the
effect of the expanding detonation products on the shape of pressure history is not significant
because the distance between the shock front and the front of the expanding detonation products
is small.
Figure 2-43 plots the arrival time of the shock front at distances of r equal to 148 mm (1 mm
inside the face of the charge) and 250 mm (101 mm from the face of the charge) as a function of
cell size. The arrival times are insensitive to the choice of cell size, of the four considered.

73

a. r = 152.7 m (3.7 mm from face of charge); Z = 0.054 m/kg1/3; r =1.025

b. r = 164 mm (15 mm from face of charge); Z = 0.058 m/kg1/3; r =1.10


Figure 2-41 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning

74

c. r = 200 mm (51 mm from face of charge); Z = 0.071 m/kg1/3; r =1.34

d. r = 250 mm (101 mm from face of charge); Z = 0.088 m/kg1/3; r =1.68


Figure 2-41 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning (cont.)

75

e. r = 500 mm (351 mm from face of charge); Z = 0.177 m/kg1/3; r =3.36

f. r = 1000 mm (851 mm from face of charge); Z = 0.353 m/kg1/3; r =6.71


Figure 2-41 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning (cont.)

76

a. r = 152.7 mm (3.7 mm from face of charge); Z = 0.054 m/kg1/3; r =1.025

b. r = 164 mm (15 mm from face of charge); Z = 0.058 m/kg1/3; r =1.10


Figure 2-42 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning

77

c. r = 200 mm (51 mm from face of charge); Z = 0.071 m/kg1/3; r =1.34

d. r = 250 mm (101 mm from face of charge); Z = 0.088 m/kg1/3; r =1.68


Figure 2-42 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning (cont.)

78

e. r = 500 mm (351 mm from face of charge); Z = 0.177 m/kg1/3; r =3.36

f. r = 1000 mm (851 mm from face of charge); Z = 0.353 m/kg1/3; r =6.71


Figure 2-42 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning (cont.)

79

Figure 2-43 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT; JWL EOS; no afterburning (cont.)
2.4.3 Temperature in the Vicinity of the Charge Face
In the vicinity of the face of the charge, the temperature is extremely high, and these high
temperatures can enable afterburning as discussed in Section 2.4.4. Temperatures for the 22.68
kg charge of TNT are monitored using a cell size of 0.5 mm at Z = 0.054, 0.058, 0.071, 0.088,
0.177 and 0.353 m/kg1/3, used in the mesh convergence study of Section 2.4.2. Afterburning is
not considered.
Histories of temperature for the range considered are presented in Figures 2-44 through 2-49,
together with those of overpressure and density. Very close to the face of the charge (i.e., Z =
0.054 m/kg1/3; Figure 2-44), the peak temperature is approximately 200000 K. The temperature
rises nearly instantaneously with the arrival of the shock front but the peak temperature occurs at
the front of the expanding detonation products. Consider Figure 2-47 (Z = 0.088 m/kg1/3). In the
density history, the first (shown with open circle) and the second peaks near t = 0.04 ms indicate
the arrival of the shock front and the front of the expanding detonation products, respectively.
The peak temperature shown in Figure 2-47c is associated with the second peak in the density
history, which is also observed in Figures 2-48 and 2-49 (Z = 0.177 and 0.353 m/kg1/3,
80

respectively). The relationship between the peak temperature and the expanding detonation
products cannot be seen for Z = 0.054, 0.058 and 0.71 m/kg1/3, as shown in Figures 2-44, 2-45
and 2-46, respectively, because the distances between the shock front and the front of the
expanding detonation products cannot be resolved at this scale.
Figure 2-50 shows peak temperature as a function of radial expansion of the detonation products
for three charges of TNT of 22.68, 960 and 18735 kg. The influence of the charge weight on the
results is insignificant. These data inform the afterburning study that is presented next.

81

a. Overpressure

b. Density

c. Temperature
Figure 2-44 Overpressure, density and temperature histories at Z = 0.054 m/kg1/3;
spherical charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh
82

a. Overpressure

b. Density

c. Temperature
Figure 2-45 Overpressure, density and temperature histories at Z = 0.058 m/kg1/3;
spherical charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh
83

a. Overpressure

b. Density

c. Temperature
Figure 2-46 Overpressure, density and temperature histories at Z = 0.071 m/kg1/3;
spherical charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh
84

a. Overpressure

b. Density

c. Temperature
Figure 2-47 Overpressure, density and temperature histories at Z = 0.088 m/kg1/3;
spherical charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh
85

a. Overpressure

b. Density

c. Temperature
Figure 2-48 Overpressure, density and temperature histories at Z = 0.177 m/kg1/3;
spherical charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh
86

a. Overpressure

b. Density

c. Temperature
Figure 2-49 Overpressure, density and temperature histories at Z = 0.353 m/kg1/3;
spherical charge of 22.68 kg of TNT; JWL EOS; 0.5 mm mesh
87

Figure 2-50 Variation of temperature with radial expansion of detonation product


2.4.4 Afterburning
To date, the analysis has assumed that all the energy of the explosive is released upon detonation
and used to drive the shock front forward. However, for under-oxidized explosives, this
assumption is not necessarily correct. For such explosives, there exists an additional energyrelease mechanism. TNT is an under-oxidized explosive that does not have sufficient oxygen to
completely oxidize. As the shock wave expands into the atmosphere, the detonation products
consume oxygen from the surrounding air and oxidize; the process is called afterburning
(Donahue 2008). Afterburning involves combustion reactions that release energy and increase
the temperature of the affected region, enhancing the effects of detonation. Unlike detonation,
where the release of energy takes place in microseconds, afterburning is a process that can last
milliseconds to seconds. There are two requirements essential for afterburning to occur: 1) the
temperature in the region should be high enough for combustion reactions to take place, and 2)
there must be enough oxygen in the ambient air. TNT is a fuel-rich explosive, with the main
88

fuels being Carbon, Carbon Monoxide, Hydrogen, and Methane. These fuels undergo
combustion reactions at their respective ignition temperatures. If the temperature is 1800+ K,
sufficient oxygen is present, and the oxygen is well mixed with the fuels, then all the combustion
reactions can occur (e.g., Souers et al. 2001, McNesby et al. 2010). The temperatures computed
by analysis exceed 10000 K during the initial stages of the expansion beyond the original surface
of the charge. Given that afterburning should occur, AUTODYN is used to model it. The
additional energy to be released due to afterburning and the duration of the release can be
specified in AUTODYN in the definition of the JWL EOS.
In Figure 2-50, the temperature decreases to 1800 K at a radial expansion of the detonation
products, rd , of 13.4 (Z = 0.71 m/kg1/3), where the radial expansion of the shock front, r , is 17.2
(Z = 0.91 m/kg1/3). For a charge weight of 22.68 kg, the volume of air between the original
surface of the charge and at a radial expansion, rd , of 13.4 is 33.3 m3. Donahue (2008) reported
that 3.18 kg of air is required per 1 kg of TNT to release the maximum afterburning energy. For
a volume of air of 33.3 m3, the maximum afterburning energy is 129 MJ (5.67 MJ/kg 22.68 kg)
(132% of the detonation energy, 96 MJ (4.23 MJ/kg 22.68 kg)). In the very early stages of the
expansion, afterburning does not significantly affect the blast wave (e.g., Cullis and HuntingtonThresher 2007, McNesby et al. 2010), which is due to a lack of available oxygen (air) and
incomplete mixing with air. Assuming that the fuels are completely mixed with the available
oxygen, the afterburning energy of 129 MJ is divided into six intervals, as noted in Table 2-14
and Figure 2-51. The front of the detonation products reaches the radial expansions, rd , of 1, 4, 6,
8, 10, 12 and 13.4 at t = 0.0215, 0.123, 0.245, 0.406, 0.610, 0.897 and 1.15 msec, respectively.
The density of air at room temperature is assumed to be 1.225 kg/m3. The maximum afterburning
energy is 10.01 MJ per 1 kg of TNT (Donahue 2008). The weight of TNT for an amount of air to
produce the maximum afterburning energy is the weight of air divided by 3.18 kg-air/kg-TNT.
The total afterburning energy is the weight of TNT multiplied by 10.01 MJ/kg-TNT. The specific
afterburning energy is thus calculated by dividing the total afterburning energy by the weight of
TNT (= 22.68 kg) and is then added to the AUTODYN analysis.

89

Table 2-14 Calculations for the afterburning energy at each interval between radial
expansions of detonation products for a spherical charge of 22.68 kg of TNT.
Range of radial

1 to 4

4 to 6

6 to 8

8 to 10

Volume of air (m3)

0.87

2.1

4.1

6.8

10.1

9.4

33.3

Weight of air (kg)

1.07

2.58

5.02

8.28

12.4

11.5

40.8

0.336

0.811

1.58

2.60

3.89

3.62

12.8

3.37

8.12

15.8

26.1

38.9

36.2

129

0.148

0.358

0.697

1.15

1.71

1.60

5.67

3.5

8.4

16.3

26.9

40.1

37.3

132

expansion,

rd

Weight of TNT for an


available air to produce
the maximum
afterburning energy
(kg)
Total afterburning
energy (MJ)
Specific afterburning
energy (MJ/kg)
Percentage of the
detonation energy

10 to 12 12 to 13

Figure 2-51 Input of afterburning energy with time for 22.68 kg of TNT

90

Total

The afterburning energy is released from a problem time of 0.0215 msec (radial expansion of
detonation products, rd = 1.0) to 1.15 msec ( rd = 13.4). Figure 2-51 presents the afterburning
energy with time, which is released at a constant rate (uniformly) in each interval of the
simulation, as noted in the third last row of Table 2-14. Figure 2-52 describes the relationship
between the weight of TNT and the available volume of air to release the maximum afterburning
energy. Figure 2-53 describes the available volume of air as a function of the radial expansion of
detonation products, rd , where the available volume of air is computed as the volume between
the face of the charge and the spherical surface corresponding to rd .
Figures 2-54 and 2-55 present the results of an analysis for a mesh of 0.5 mm cells, assuming a
total afterburning energy of 129 MJ (132% of the detonation energy; see Figure 2-51). For this
explosive, the effect of afterburning (as modeled) is insignificant at a distance of 200 mm
(corresponding to rd = 1.34) due to a lack of available oxygen, but an increase in impulse is
evident at the greater distance of 250 mm ( rd = 1.68). This increase is an upper bound on the
effect of afterburning because perfect mixing of the available fuel and oxygen is assumed.

91

Figure 2-52 Weight of TNT for volumes of air to release maximum afterburning energy

Figure 2-53 Available volumes of air for radial expansion of detonation products

92

a. r = 200 mm (51 mm from face of charge)

b. r = 250 mm (101 mm from face of charge)


Figure 2-54 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT for a 0.5 mm mesh; JWL EOS; 129 MJ afterburning
93

c. r = 500 mm (351 mm from face of charge)

d. r = 1000 mm (851 mm from face of charge)


Figure 2-54 Overpressure histories at distance, r, from the center of a spherical charge of
22.68 kg of TNT for a 0.5 mm mesh; JWL EOS; 129 MJ afterburning (cont.)
94

a. r = 200 mm (51 mm from face of charge)

b. r = 250 mm (101 mm from face of charge)


Figure 2-55 Impulse histories at distance, r, from the center of a spherical charge of 22.68
kg of TNT for a 0.5 mm mesh; JWL EOS; 129 MJ afterburning
95

c. r = 500 mm (351 mm from face of charge)

d. r = 1000 mm (851 mm from face of charge)


Figure 2-55 Impulse histories at distance, r, from the center of a spherical charge of 22.68
kg of TNT for a 0.5 mm mesh; JWL EOS; 129 MJ afterburning (cont.)
96

2.4.5 Blast Wave Propagation in AUTODYN and Air3D


One-dimensional blast wave propagation is studied using Air3D to contrast the predictions of
Air3D and AUTODYN. Air3D (Rose 2006) is a computational fluid dynamics (CFD) code
developed specifically for blast applications. The input data is specified via a text file, which is
then run in batch mode. In a 3D analysis, there are three sections of the input file that are
executed sequentially: spherical (1D), radial (2D) and main (3D). When the spherical input
section is executed, the calculations proceed with spherical symmetry until the blast wave
reaches a reflecting surface. The output of the 1D analysis is then remapped into 2D (the x, y
domain) and then to a 3D domain.
Air3D uses the balloon analog of Ritzel and Matthews (1997) and assumes a constant for air.
There are two versions of Air3D: ideal gas EOS and JWL EOS. The ideal gas version is used
here. The input needed for the 1D section of the input file is initial weight (= 22.68 kg), density
(= 1637 kg/m3) and energy (= 4.3106 J/kg) of the explosive, the cell size for the 1D calculations
(= 0.5 mm) and the radius for the 1D calculations (= 2 m). In the 2D input section, the radial and
axial boundaries for the 2D domain are specified by rmax (= 3 m) and hmax (= 3 m). The
boundary conditions are specified by bru (-1) and hru (+1) in the input file; -1 represents a
stop boundary and +1 represents a transmit boundary. The Air3D User Guide (Rose 2006)
recommends a maximum cell size based on scaled distance; i.e., the cell size should represent
roughly a scaled radial distance of 1 mm/kg1/3. If the mass of the explosive is 22.68 kg, the cell
size should be less than 122.681/3 or 2.82 mm. For a packing density of 1637 kg/m3 and a
charge radius of 149 mm, the initial pressure in the balloon is 2816 MPa.
Figure 2-56 presents fringe plots of pressure from the 1D AUTODYN analysis of the detonation
of a 22.68 kg charge of TNT at time instants after detonation. The location of the shock front is
identified by a fine line that is initially at a distance r = 149 mm at t = 0.015 and 0.020 msec, and
r = 480 mm at t = 0.1 msec. The detonation front reached the face of the charge at a time
between 0.020 msec and 0.025 msec. The shock wave then propagates continuously outward
from the face of the charge and a rarefaction wave begins to propagate toward the center of the
charge. Before the head of the rarefaction wave reaches the center of the charge, the peak

97

pressure at the center is approximately 4.2 GPa. The head of the rarefaction wave arrives at the
center of the charge near t = 0.060 msec and the pressure at the center decreases rapidly.
Figure 2-57 presents results of the 1D Air3D analysis for the same problem used in the
AUTODYN analysis. The assumed detonation wave speed for the TNT is 6730 m/second (Rose
2006) and so the starting time for analysis is 0.02214 msec (= 0.149/6730) and shortly following
the time associated with the first panel in the figure (= 0.022 msec). The initial pressure of the
gas in the balloon is 2816 MPa, as noted previously. At t = 0.095 msec, the head of the
rarefaction wave has just reached the point of detonation. The shock front is clearly observed at
t = 0.146 msec, at which the tail of the rarefaction wave is near the front of the detonation
products. The region between the shock front and the tail of the rarefaction wave is compressing,
whereas the region behind the tail of the rarefaction wave is expanding. The pressures behind the
tail of the rarefaction wave continuously decrease and become small enough to ignore at t =
0.500 msec.

98

Time
(msec)

Pressure fringes

0.015

0.020

0.025

0.030

Figure 2-56 Overpressure histories as a function of time after detonation at distance x (= r)


from the center of a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh;
JWL EOS; no afterburning

99

Time
(msec)

Pressure fringes

0.035

0.040

0.045

0.050

Figure 2-56 Overpressure histories as a function of time after detonation at distance x (= r)


from the center of a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh;
JWL EOS; no afterburning (cont.)

100

Time
(msec)

Pressure fringes

0.055

0.060

0.065

0.070

Figure 2-56 Overpressure histories as a function of time after detonation at distance x (= r)


from the center of a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh;
JWL EOS; no afterburning (cont.)

101

Time
(msec)

Pressure fringes

0.075

0.080

0.085

0.090

Figure 2-56 Overpressure histories as a function of time after detonation at distance x (= r)


from the center of a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh;
JWL EOS; no afterburning (cont.)

102

Time
(msec)

Pressure fringes

0.095

0.100

Figure 2-56 Overpressure histories as a function of time after detonation at distance x (= r)


from the center of a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh;
JWL EOS; no afterburning (cont.)

103

Time (msec)

Overpressure with distance

0.022

0.059

0.095

Figure 2-57 Air3D overpressure histories as a function of distance and time after
detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh; ideal
gas EOS

104

Time (msec)

Overpressure with distance

0.103

0.111

0.128

Figure 2-57 Air3D overpressure histories as a function of distance and time after
detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh; ideal
gas EOS (cont.)

105

Time (msec)

Overpressure with distance

0.146

0.173

0.229

Figure 2-57 Air3D overpressure histories as a function of distance and time after
detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh; ideal
gas EOS (cont.)

106

Time (msec)

Overpressure with distance

0.287

0.392

0.493

Figure 2-57 Air3D overpressure histories as a function of distance and time after
detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh; ideal
gas EOS (cont.)

107

2.4.6 Equations of State


To evaluate the ideal gas EOS in Air3D for modeling near-field detonations, the simulation
results are compared with those using both the ideal gas EOS and the JWL EOS implemented in
AUTODYN. Figure 2-58 presents the results of 1D analysis using a) AUTODYN, JWL EOS for
the explosive and the ideal gas EOS for air, b) AUTODYN and the ideal gas EOS ( 1.4 ) for
both explosive and air, and c) Air3D and the ideal gas EOS ( 1.4 ) for explosive and air.
Results are presented for r = 148 mm, 1 mm inside the face of the charge (panel a), r = 200 mm
(panel b), r = 250 mm (panel c), r = 500 mm (panel d), and r = 1000 mm (panel e). Significant
differences are observed for each value of r between the JWL EOS and the ideal gas EOS. At 1
mm inside the face of the charge, there are differences between all three models: the AUTODYN
ideal gas EOS incident peak overpressure is dramatically smaller than that predicted using the
JWL EOS, and the use of the Air3D balloon analog results in a higher incident overpressure at
the face of the charge than AUTODYN using the ideal gas EOS. At r = 200 mm (radial
expansion of 1.34, volume expansion of 2.42), the AUTODYN and Air3D calculations using the
ideal gas EOS provide similar predictions of peak pressure associated with the shock front (20
MPa) and the expanding detonation products (120 MPa). Both ideal gas computations
underpredict the incident peak overpressure of the JWL EOS (40 MPa) and predict a greater
value of peak overpressure associated with the expanding detonation products and a greater
incident impulse. Note that the scale of overpressure in panel a is approximately two orders of
magnitude higher than in panel b. An analysis of panel c (r = 250 mm, radial expansion = 1.68,
volume expansion = 4.72) leads to similar observations. At a distance of r = 500 mm (351 mm
from the face of the charge, radial expansion = 3.36, volume expansion = 37.8), the AUTODYN
JWL computations predict a greater value of incident peak overpressure than the ideal gas
computations. (The AUTODYN implementation of the JWL EOS collapses to an ideal gas EOS,
default 1.3 , at a volume expansion = 10, radial expansion = 2.15.) The differences between
the three overpressure histories and incident impulses at r = 1000 mm (radial expansion = 6.71,
scaled range = 0.353 m/kg1/3) are relatively small.
The ideal gas EOS should not be used for air-blast calculations at small values of scaled distance,
say r 7 based on the simulation results presented here.
108

a. r = 0.148 m (1 mm inside face of charge)

b. r = 0.200 m (51 mm from face of charge)


Figure 2-58 Influence of EOS on overpressure histories as a function of distance and time
after detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh

109

c. r = 0.250 m (101 mm from face of charge)

d. r = 0.500 m (351 mm from face of charge)


Figure 2-58 Influence of EOS on overpressure histories as a function of distance and time
after detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh
(cont.)
110

e. r = 1.000 m (851 mm from face of charge)


Figure 2-58 Influence of EOS on overpressure histories as a function of distance and time
after detonation for a spherical charge of 22.68 kg of TNT for a 0.5 mm mesh
(cont.)

2.5 Two-Dimensional Blast Wave Propagation


Two-dimensional blast wave propagation is simulated to predict reflected overpressures in the
near field. The effects of boundary conditions, optimal cell sizes in 2D, consistency between 1D
and 2D, and near-field effects on reflected overpressure histories are investigated.
Three 2D models are evaluated to examine the effect of boundary conditions on overpressure and
impulse histories. The first two models include one quarter of the charge and a varied vertical
boundary condition at the centerline of the (spherical) charge. The third model includes one half
of the charge and places a transmitting (TR) vertical boundary away from the charge. Each
model is described below.
Consider first the 2D Model I-I shown in Figure 2-59a. A quarter of the charge is modeled in a
2D domain with dimensions of 300200 mm (horizontal vertical). The spherical charge is
22.68 kg of TNT with a packing density of 1630 kg/m3 and a charge radius of 149 mm. The left
111

vertical boundary condition in this model is set as transmitting. The other two models investigate
the use of alternate left vertical boundary conditions.
Figure 2-59 examines the effect of the left vertical boundary in Model I-I on incident
overpressure and impulse histories. Energy should not escape across a left transmitting boundary
because the pressure wave initially propagates radially from the source and away from this
boundary. The left vertical TR boundary is replaced by a (mirrored-symmetric) reflecting (RE)
boundary. Incident overpressure histories are calculated at monitoring points 1, 2 and 3 (see
Figure 2-59a) using both left TR and RE boundaries. The coordinates of monitoring points 1, 2
and 3 are at (x, y) = (200, 0), (0, 80) and (0, 160), respectively. A cell size of 0.5 mm is used for
the analysis because the 1D analysis results converged at a distance of 200 mm for the 0.5 mm
cell size (see Figure 2-41c). Overpressure and impulse histories are presented in Figures 2-59b,
2-59c and 2-59d at the three monitoring points. The histories are identical at each monitoring
point. The differences are less than 1%.

a. Model I-I
Figure 2-59 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-I

112

b. Overpressure and impulse histories at monitoring location 1

c. Overpressure and impulse histories at monitoring location 2

d. Overpressure and impulse histories at monitoring location 3


Figure 2-59 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-I (cont.)

113

Figure 2-60 enables a comparison of results of 1D (radial) and 2D analysis for the 2D Model I-I
with the left transmitting boundary. Results of the two analyses are presented for a 0.5 mm mesh
at monitoring location 1. The overpressure histories are identical, which is not surprising given
the use of radial (axial) symmetry in both cases.

Figure 2-60 AUTODYN analysis results for 1D and 2D (I-I with the left transmitting
boundary) models; spherical charge of 22.68 kg of TNT; 0.5 mm mesh;
monitoring location 1;
Another 2D model, I-II, is presented in Figure 2-61a. One half of the spherical charge is modeled
using a lower axially symmetric boundary. The left transmitting boundary of Model I-I is moved
to the left at distance of two charge radii. The monitoring locations are same as those used in
Model I-I. One might argue that this model is better than Model I-I with a left vertical
transmitting boundary but it is computationally more expensive. Figures 2-61b, 2-61c and 2-61d
enable a comparison of overpressure and impulse histories calculated using 1) Model I-I with a
left vertical transmitting boundary, and 2) Model I-II, at monitoring locations, 1, 2 and 3,
respectively. The histories are identical for all three monitoring locations.
Models I-III and I-IV, as shown in Figures 2-62a and 2-62b, respectively, are constructed from
Models I-I and I-II, respectively, to study the effect of choice of boundary condition on reflected
overpressure histories. The right vertical transmitting boundaries of Figures 2-59a and 2-61a are
replaced by reflecting boundaries. Overpressure and impulse histories are calculated at five
monitoring locations, where monitoring locations 1, 2 and 3 were identified previously and

114

monitoring locations 4, 5 and 6 are on the right reflecting boundary, 0, 80 and 160 mm,
respectively, from the lower horizontal boundary.
Figures 2-62c through 2-62n enable a comparison of the overpressure and impulse histories
calculated using models I-III and I-IV. In the legends, TR and RE identify the left transmitting
and reflecting boundaries, respectively, for model I-III. The overpressure histories are plotted at
two ranges on the y (overpressure) axis, full and part, where the part-range plots enable analysis
of small differences in the histories. Analysis using Model I-III TR and Model I-IV produce
identical results. The effects of multiple reflections from the model with two vertical reflecting
boundaries (Model I-III RE) are clearly seen in the impulse histories at monitoring locations 1, 3,
4, 5 and 6.
In summary, analysis of a model that incorporates a left vertical transmitting boundary at the
centerline of the charge produces identical overpressures and impulses to analysis of a model of
one-half of the charge with a left transmitting boundary placed two charge radii from the center
of the weapon. Hereafter, computationally more efficient models involving a vertical
transmitting boundary at the center of the charges are analyzed.

a. Model I-II
Figure 2-61 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-II

115

b. Overpressure and impulse histories at monitoring location 1

c. Overpressure and impulse histories at monitoring location 2

d. Overpressure and impulse histories at monitoring location 3


Figure 2-61 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-II (cont.)
116

a. Model I-III

b. Model I-IV

c. Overpressure histories in full (left) and part (right) ranges at monitoring location 1
Figure 2-62 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-III and Model I-IV

117

d. Impulse history at monitoring location 1

e. Overpressure histories in full (left) and part (right) ranges at monitoring location 2

f. Impulse history at monitoring location 2


Figure 2-62 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-III and Model I-IV (cont.)
118

g. Overpressure histories in full (left) and part (right) ranges at monitoring location 3

h. Impulse history at monitoring location 3

i. Overpressure histories in full (left) and part (right) ranges at monitoring location 4
Figure 2-62 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-III and Model I-IV (cont.)
119

j. Impulse history at monitoring location 4

k. Overpressure histories in full (left) and part (right) ranges at monitoring location 5

l. Impulse history at monitoring location 5


Figure 2-62 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-III and Model I-IV (cont.)
120

m. Overpressure histories in full (left) and part (right) ranges at monitoring location 6

n. Impulse history at monitoring location 6


Figure 2-62 Effect of boundary condition; spherical charge of TNT of 22.68 kg; 0.5 mm
cells; Model I-III and Model I-IV (cont.)
Figure 2-63 presents the results of 2D AUTODYN analysis with Model II, as shown in panel a.
The radius of the charge is 149 mm. For this 2D analysis, 1D results using a cell size of 0.5 mm,
chosen for the same reason described previously, are remapped into 2D. Cell sizes of 4, 2, 1 and
0.5 mm are used in the 2D analyses. The TNT packing density is 1636.8 kg/m3, resulting in a
charge weight of 22.68 kg (50.00 lb). The JWL EOS is used for analysis. Afterburning is ignored
because the afterburning energy is less than 5% of the detonation energy due to the availability
of oxygen. Figure 2-63a presents the AUTODYN model, the remaining boundary conditions and
the locations of monitoring locations 2, 3 and 4 on the right hand boundary. The other panels in
the figure present reflected overpressure histories at monitoring locations 2, 3 and 4. Monitoring
locations 2, 3 and 4 are associated with angles of incidence of 0, 36.7 and 56.1, respectively,
121

and Z = 0.071, 0.088 and 0.13 m/kg1/3. The results presented at monitoring locations 2, 3 and 4
are essentially identical for all four cell sizes; the effect of cell size on the results is insignificant.
The jagged nature of the overpressure history at monitoring location 4, which may be
counterintuitive, is studied using the pressure fringes presented in Figure 2-64. Two ranges of
pressure are provided: full range (capturing the entire range of pressure) and fixed range
(intended to capture the propagation of the shock front and the expanding detonation products).
It is evident that the pressure histories at the monitoring locations are complex, which is due
primarily to the expanding detonation products. At 0.029 msec, the shock front impinges on the
vertical reflecting surface; the initial reflection of the shock front is captured in the panels at
0.030 msec, noting that the reflected pressure at monitoring location 2 increases to a maximum
value of 2500 MPa at a later time. The presence of the expanding detonation products is
observed immediately behind the shock front, as shown in the fixed range (second column) of
Figure 2-64, but the distance between the front of the expanding detonation products and the
shock front is so small that the arrival of the expanding detonation products cannot be identified
in the overpressure histories at monitoring location 2 (see Figure 2-63b) at the scale presented.
At 0.037 msec, the shock front impinges on the reflecting surface at monitoring location 3. The
arrival of the expanding detonation products is not seen in the overpressure histories at
monitoring location 3 (see Figure 2-63c) but it is observed behind the shock front in the fixed
range of Figure 2-64. At 0.054 msec (fourth row of Figure 2-64), the incident shock wave has
reached monitoring location 4 and the Mach stem has formed (between monitoring locations 2
and 3). At monitoring location 4, the arrivals of the shock front and of the front of the expanding
detonation products can be distinguished in the overpressure histories of Figure 2-63d at
approximately t = 0.5 ms and 0.7 ms, respectively. The pressure fringes immediately behind the
Mach stem are complex and irregular and provide an explanation for the jagged nature of the
pressure history seen in Figure 2-63d.

122

Transmitting boundary

44

Radius
of TNT

33

Reflecting boundary

Transmitting boundary

149 mm

149 mm 447 mm

149 mm
Detonation
point
O

Axially symmetric
boundary
149 mm
51 mm
200 mm

22

a. Model II

b. Monitoring location 2
Figure 2-63 Reflected overpressure histories on a vertical reflecting surface; Model II;
spherical charge of 22.68 kg of TNT; JWL EOS; varying mesh size

123

c. Monitoring location 3

d. Monitoring location 4
Figure 2-63 Reflected overpressure histories on a vertical reflecting surface; Model II;
spherical charge of 22.68 kg of TNT; JWL EOS; varying mesh size (cont.)

124

Time
(msec)

Pressure fringes
Full range

Fixed range

0.029

0.030

Figure 2-64 Pressure fringes as a function of time; Model I; spherical charge of 22.68 kg of
TNT; JWL EOS; TNT density = 1636.8 kg/m3; 1.0 mm mesh

125

Time
(msec)

Pressure fringes
Full range

Fixed range

0.037

0.054

Figure 2-64 Pressure fringes as a function of time; Model I; spherical charge of 22.68 kg of
TNT; JWL EOS; TNT density = 1636.8 kg/m3; 1.0 mm mesh (cont.)

126

Time
(msec)

Pressure fringes
Full range

Fixed range

0.064

Figure 2-64 Pressure fringes as a function of time; Model I; spherical charge of 22.68 kg of
TNT; JWL EOS; TNT density = 1636.8 kg/m3; 1.0 mm mesh (cont.)
Figure 2-65a presents Model III that is similar to Model II (Figure 2-63) except that the upper
boundary is extended to a height of 1500 mm above the detonation point. Results are presented
at three additional monitoring locations on the vertical reflecting surface: 5 (500 mm above the
detonation point, panel b, angle of incidence = 68.2), 6 (1000 mm above the detonation point,
panel c, angle of incidence = 78.7), and 7 (1500 mm above the detonation point, panel d, angle
of incidence = 82.4). Two cell sizes are used to generate results at monitoring locations 5 and 6;
namely, 1 mm and 0.5 mm. The reflected pressure history at monitoring location 7 is computed
using 1 mm cells. The jagged pressure histories are a result of wave reflection and Mach stem
formation, as noted previously.

127

Reflecting boundary

Transmitting boundary

Transmitting boundary
G7

F6
1500 mm
1000
mm

E5
D4

Detonation
point
O

C3

500
mm

B2
200 mm Axially symmetric
boundary

a. Model III

b. Monitoring location 5
Figure 2-65 Reflected overpressure histories on a vertical reflecting surface; Model III;
spherical charge of 22.68 kg of TNT; JWL EOS; varying mesh size

128

c. Monitoring location 6

d. Monitoring location 7
Figure 2-65 Reflected overpressure histories on a vertical reflecting surface; Model III;
spherical charge of 22.68 kg of TNT; JWL EOS; varying mesh size (cont.)

129

2.6 Three-Dimensional Wave Propagation


Three-dimensional blast wave simulations enable evaluation of reflected overpressure histories
in complex environments with multiple reflecting boundaries (e.g., explosion in free air with
buildings and ground).
Three-dimensional meshes of Model II are constructed with different mesh sizes and different
values of the perpendicular (y) dimension, noting that the plane of Model II (and IV) shown in
Figure 2-63a is x (horizontal) and z (vertical). Model II data are used because 2D axial symmetry
results are available to verify the 3D models. The 3D variant of Model II denoted as Model IV
with y = 200 mm is shown in Figure 2-66a. The vertical planes through the center of the
explosive are modeled as transmitting and reflecting boundaries in the yz and xz planes,
respectively. Similar to the 2D analyses of Section 2.5, results of 1D analyses performed using
the 0.5 mm cell size are remapped into the 3D meshes. Afterburning is not modeled for the
reason given in Section 2.5. Figure 2-66 presents reflected overpressure histories at monitoring
locations 2, 3 and 4, located 200 mm from the centerline of the charge at x = 200, 200 and 200
mm; z = 0, 149 and 298 mm; and varying y. Because of concerns regarding the transmissibility
of the boundary in the perpendicular (y) direction if the dimension was small, six values of the y
dimension in Model IV are studied: 8, 12, 48, 100, 149 and 200 mm. Computational cells of 4
mm (4.027 mm for y = 149 mm) are used to construct each mesh. Analysis of the three panels b,
c and d in Figure 2-66 makes the influence of the choice of perpendicular dimension clear, with
results stabilizing for a value of 48 mm (R/4) and greater for peak overpressure and a value for
149 mm (3R/4) and greater for impulse, where R is the shortest distance from the point of
detonation (center of the charge) to the vertical reflecting surface (= 200 mm in this case),
namely, the standoff distance. The increase in overpressure at 0.2 msec at each monitoring
location for the 100 mm perpendicular dimension (y) is due to reflections from the transmitting
perpendicular boundary. Figure 2-67 (2-68) presents the results of an analysis of mesh size for y
= 100 mm (200 mm). Results of 2D analysis invoking radial (axial) symmetry are presented to
inform the choice of cells size. For a given perpendicular dimension, 100 mm or 200 mm, the
overpressure histories are identical for the 2 mm and 4 mm cells at monitoring location 3 and 4;
the reflected peak overpressure is underestimated by approximately 10% at monitoring location 2
130

with the 4 mm cells. Both 3D meshes overestimate the reflected overpressure at 0.2 msec for y =
100 mm. Figure 2-69 enables conclusions to be drawn regarding cell size and the minimum
required perpendicular dimension for 3D analysis. Each panel in Figure 2-69 presents
overpressure histories at a given monitoring location for 2D radial analysis (2 mm mesh) and 3D
analysis (2 mm mesh) for perpendicular (y) dimensions of 100 mm and 200 mm. The histories
for 2D analysis (2 mm mesh) and 3D analysis (2 mm mesh, y = 200 mm) are essentially identical.
Transmitting boundaries do not function perfectly and a fraction of the pressure is reflected,
causing both the overpressures and impulses to be over-predicted if the transverse (y) dimension
is too small. On the basis of this study, the transverse dimension should be always specified to be
greater than the charge radius to avoid significant boundary effects.
Figure 2-70 presents Model V that is similar to Models II and IV except that the lower boundary
is extended down to the ground and one-fourth of the explosive is modeled. The lower boundary
is modeled as a reflecting surface. The vertical planes through the center of the explosive are
modeled as transmitting and reflecting boundaries in the yz and xz planes, respectively. Twodimensional radial (axial) symmetry cannot be used to solve this problem and three-dimensional
models must be used. Alternate three-dimensional models are considered for the purpose of
monitoring overpressure histories on the centerline of the charge at monitoring locations 1, 2, 3
and 4 per Figure 2-66.
On the basis of the data presented in Figures 2-66 through 2-69, Model V is meshed with 1, 2
and 4 mm cells and a perpendicular (y) dimension of 200 mm and uses 1D results for a cell size
of 0.5 mm and a distance of 149 mm, which is adopted because 1D simulations should be
performed before reflection occurs and the shock front is first reflected at the distance of 149 mm,
on the lower reflecting boundary. Figure 2-71 presents reflected overpressure histories at
monitoring locations 1, 2, 3 and 4, with all data presented in panel a (4 mm mesh) and histories
for individual monitoring locations presented in panels b, c, d and e. The results are effectively
identical for the cell sizes of 1, 2 and 4 mm. Results for the 1 mm cell size are not included at
monitoring location 4 because the runtimes were excessive. The reflected overpressure at the
vertical reflecting surface at the ground is 3.8 GPa (2 mm mesh) and approximately twice that at
monitoring location 2 (2.1 GPa, 2 mm), 149 mm above monitoring location 1. The reflected peak
131

overpressure at monitoring location 2 calculated using Model V and a 2 mm mesh is


approximately 10 percent less than that calculated using Model II, axial symmetry and a 1 mm
mesh. The reflected peak overpressure at monitoring location 3 is similar for Model II (1 mm
mesh) and Model V (2 mm mesh). The local increase in the overpressure history at monitoring
location 2 at t = 0.15 msec is due to the formation of a Mach stem. None of the Model V pressure
histories resembles the shape of the Friedlander equation.

a. Model IV; 3D variant of Model II; perpendicular (y) dimension = 200 mm


Figure 2-66 Reflected overpressure histories on a vertical reflecting surface; Model IV;
spherical charge of 22.68 kg of TNT; JWL EOS; 4 mm mesh (4.027 mm for y =
149 mm); varying perpendicular dimension, y

132

b. Monitoring location 2

c. Monitoring location 3
Figure 2-66 Reflected overpressure histories on a vertical reflecting surface; Model IV;
spherical charge of 22.68 kg of TNT; JWL EOS; 4 mm mesh (4.027 mm for y =
149 mm); varying perpendicular dimension, y (cont.)

133

d. Monitoring location 4
Figure 2-66 Reflected overpressure histories on a vertical reflecting surface; Model IV;
spherical charge of 22.68 kg of TNT; JWL EOS; 4 mm mesh (4.027 mm for y =
149 mm); varying perpendicular dimension, y (cont.)

134

a. Monitoring location 2

b. Monitoring location 3
Figure 2-67 Reflected overpressure histories on a vertical reflecting surface; Model II and
Model IV; spherical charge of 22.68 kg of TNT; JWL EOS; y = 100 mm for 3D
analysis

135

c. Monitoring location 4
Figure 2-67 Reflected overpressure histories on a vertical reflecting surface; Model II and
Model IV; spherical charge of 22.68 kg of TNT; JWL EOS; y = 100 mm for 3D
analysis (cont.)

136

a. Monitoring location 2

b. Monitoring location 3
Figure 2-68 Reflected overpressure histories on a vertical reflecting surface; Model II and
Model IV; spherical charge of 22.68 kg of TNT; JWL EOS; y = 200 mm for 3D
analysis

137

c. Monitoring location 4
Figure 2-68 Reflected overpressure histories on a vertical reflecting surface; Model II and
Model IV; spherical charge of 22.68 kg of TNT; JWL EOS; y = 200 mm for 3D
analysis (cont.)

138

a. Monitoring location 2

b. Monitoring location 3
Figure 2-69 Reflected overpressure histories on a vertical reflecting surface; Model II and
Model IV; spherical charge of 22.68 kg of TNT; JWL EOS; 2 mm mesh

139

c. Monitoring location 4
Figure 2-69 Reflected overpressure histories on a vertical reflecting surface; Model II and
Model IV; spherical charge of 22.68 kg of TNT; JWL EOS; 2 mm mesh (cont.)

140

Figure 2-70 Reflected overpressure histories on a vertical reflecting surface; Model V;


spherical charge of 22.68 kg of TNT; JWL EOS; y = 200 mm

141

a. All monitoring locations (4 mm mesh)

b. Monitoring location 1
Figure 2-71 Reflected overpressure histories on a vertical reflecting surface; Model V;
spherical charge of 22.68 kg of TNT; JWL EOS; 2 mm and 4 mm meshes

142

c. Monitoring location 2

d. Monitoring location 3
Figure 2-71 Reflected overpressure histories on a vertical reflecting surface; Model V;
spherical charge of 22.68 kg of TNT; JWL EOS; 2 mm and 4 mm meshes
(cont.)
143

e. Monitoring location 4
Figure 2-71 Reflected overpressure histories on a vertical reflecting surface; Model V;
spherical charge of 22.68 kg of TNT; JWL EOS; 2 mm and 4 mm mesh meshes
(cont.)

2.7 Air-Blast Loadings on Sample Reinforced Concrete Columns


2.7.1 Introduction
The studies of the prior sections are extended here to the calculation of near-field air-blast
loadings on the perimeters of a square column and a circular column similar in size to those used
in buildings and mission-critical infrastructure. The goals are to compare and contrast the
reflected overpressures as a function of monitoring location on each column, and to establish the
importance of column shape on net air-blast loadings.
The dimensions of the square column are 1000 mm 1000 mm. The diameter of the circular
column is 1000 mm. Both columns are assumed rigid for the purpose of the analysis. The weight
of the TNT charge is 22.68 kg; the packing density of the TNT is 1636.8 m/kg1/3, and the radius
of the charge is 149 mm. The shortest clear distance between the face of the charge and each
column is 51 mm per Models II, III, IV and V. The centerline of each column aligns with the

144

centerline of the charge. The spherical charge is placed on a rigid reflecting surface. Symmetry is
invoked on the centerline of the charge and column. Air is modeled using an Eulerian grid and
the columns are modeled using a Lagrangian grid. The Lagrangian grid serves as a flow
constraint for the Eulerian grid.
2.7.2 Square Column, Model VI
Figure 2-72 presents a plan view and an elevation of Model VI. The x-y-z coordinate system is
shown. The transmitting y-boundary is set at y = 700 mm: 200 mm beyond the face of the
column. The transmitting z-boundary is set at z = 600 mm because reflected overpressures were
expected to be a factor of 25+ smaller at z = 600 than z = 0 on the centerline of the column.
Figure 2-73 presents the monitoring locations. Monitoring locations 1, 2, 3 and 4 are evenly
spaced at 160 mm across the front face of the column. Monitoring locations 5, 6, 7 and 8 are
evenly spaced at 250 mm along the perpendicular face of the column. Monitoring location 10
aligns with Monitoring location 1 and Monitoring location 9 is midway (250 mm) between
monitoring locations 8 and 10. Monitoring locations 11 through 20, 21 through 30, and 31
through 40 align vertically with Monitoring locations 1 through 10, respectively, but are 149 mm,
298 mm and 447 mm above the ground surface, respectively. A cell size of 4 mm is used based
on the 3D analysis of Model V presented in Section 2.6.
Figure 2-74 presents reflected overpressure histories at all 40 monitoring locations. As expected,
the reflected overpressures are greatest on the front face and (relatively) insignificant elsewhere.
A comparison of the reflected peak overpressure histories at monitoring locations 11 (opposite to
the center of the charge) and 34 [31], on the front face of the column, shows a difference of a
factor greater than 50 [15], which is due to differences in scaled distance and angle of incidence.
Figure 2-75 presents net loading and impulse histories on the lower 600 mm of the front face of
the column calculated using a) overpressure history at locations 1, 11, 21 and 31, applied across
the width of the column using tributary heights, and b) overpressure history at all monitoring
locations on the front face, applied using tributary areas. The peak load (impulse) calculated
considering all locations is 3.7 (2.3) times smaller than that calculated based on pressure histories
at locations 1, 11, 21 and 31 only. At a distance of 200 mm (Z = 0.071 m/kg1/3), discretization of
145

the front face of a square column will result in significantly smaller net loads (73% reduction
here) and impulses (57% reduction here).
Panels a through d of Figure 2-76 enable a comparison of the overpressure histories from Model
V (Figure 2-71a, locations 1, 2, 3 and 4) and Model VI (Figure 2-74i, locations 1, 11, 21 and 31),
noting that locations 1, 2, 3 and 4 in Model V are identical to 1, 11, 21 and 31 in Model VI. The
two models yield identical results.
Transmitting surface

500 mm

Detonation point

Transmitting surface

Transmitting surface

200 mm

Reflecting surface
Column

Reflecting surface
200 mm

1000 mm
1400 mm

200 mm
x

a Plan view

Column

Transmitting surface

600 mm

Transmitting surface

Transmitting surface

Detonation point

Reflecting surface
1400 mm

x
y dfdsfds

b. Elevation view
Figure 2-72 Model VI; spherical charge of 22.68 kg of TNT; square column 10001000 mm

146

a. View 1

b. View 2
Figure 2-73 Monitoring locations for Model VI; spherical charge of 22.68 kg of TNT;
square column 10001000 mm

147

a. Monitoring locations 1, 2, 3 and 4

b. Monitoring locations 5, 6, 7, 8, 9 and 10


Figure 2-74 Reflected overpressure histories; Model VI; spherical charge of 22.68 kg of
TNT; JWL EOS

148

c. Monitoring locations 11, 12, 13 and 14

d. Monitoring locations 15, 16, 17, 18, 19 and 20


Figure 2-74 Reflected overpressure histories; Model VI; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

149

e. Monitoring locations 21, 22, 23 and 24

f. Monitoring locations 25, 26, 27, 28, 29 and 30


Figure 2-74 Reflected overpressure histories; Model VI; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

150

g. Monitoring locations 31, 32, 33 and 34

h. Monitoring locations 35, 36, 37, 38, 39 and 40


Figure 2-74 Reflected overpressure histories; Model VI; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

151

i. Monitoring locations 1, 11, 21 and 31


Figure 2-74 Reflected overpressure histories; Model VI; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

152

a. Net loading

b. Impulse history
Figure 2-75 Net loading and impulse histories; Model VI

153

a. Monitoring locations 1 (Model V) and 1 (Model VI)

b. Monitoring locations 2 (Model V) and 11 (Model VI)


Figure 2-76 Comparison of reflected overpressure histories from Models V and VI

154

c. Monitoring locations 3 (Model V) and 21 (Model VI)

d. Monitoring locations 4 (Model V) and 31 (Model VI)


Figure 2-76 Comparison of reflected overpressure histories from Models V and VI (cont.)

155

2.7.3 Circular Column, Model VII


Figure 2-77 presents a plan view and an elevation of Model VII. The x-y-z coordinate system is
shown. The transmitting y-boundary is set at y = 700 mm: 200 mm beyond the face of the
column. The transmitting z-boundary is set at z = 600 mm. Figure 2-78 presents the monitoring
locations. There are eight monitoring locations around the half-perimeter of the column. The y
coordinates of monitoring locations 1 through 8 are 0, 160, 320, 480, 480, 320, 160 and 0 mm.
Monitoring locations 9 through 16, 17 through 24, and 25 through 32 align vertically with
monitoring locations 1 through 8, respectively, but are 149 mm, 298 mm and 447 mm above the
ground surface, respectively. A cell size of 4 mm is used for the analysis per Section 2.7.2.
Figure 2-79 presents reflected pressure histories at all 32 monitoring locations. The pressures are
greatest on the lines 1 through 25 and 2 through 26 (see Figure 2-78a), and (relatively)
insignificant elsewhere, as expected. A comparison of the reflected peak overpressure histories at
monitoring locations 9 (opposite to the center of the charge) and 28 [25], on the front face of the
column, shows a difference of a factor greater than 245 [13], which is due to differences in
scaled distance and angle of incidence.
Figure 2-80 presents net loading and impulse histories on the front face of the column calculated
using a) overpressure history at locations 1, 9, 17 and 25, applied across the width of the column
using tributary heights, and b) overpressure history at all monitoring locations on the front face,
applied using tributary areas. For the calculations of the net loading and impulse histories for the
circular column, Model VII, only translational (x) loads in the direction of blast waves are
considered. The peak load (impulse) calculated considering all locations is 4.7 (3.5) times
smaller than that calculated based on pressure histories at locations 1, 9, 17 and 25 only. Similar
to the analysis of the square column, Model VI, and at the scaled distance of 0.071 m/kg1/3,
discretization of the front face of a circular column will result in significantly smaller net loads
(79% reduction here) and impulses (72% reduction here). These differences in the reflected
overpressures and impulses are more substantial for the circular shape of column than that of the
square shape.

156

Transmitting surface

Transmitting surface

Transmitting surface

700 mm

Detonation point

Reflecting surface
200 mm

1000 mm

200 mm

1400 mm

a. Plan view

Transmitting surface

600 mm

Transmitting surface

Transmitting surface

Detonation point

Reflecting surface
1400 mm

dfdsfds

b. Elevation view
Figure 2-77 Model VII; spherical charge of 22.68 kg of TNT; 1000 mm diameter column

157

a. View 1

b. View 2
Figure 2-78 Monitoring locations for Model VII; spherical charge of 22.68 kg of TNT;
1000 mm diameter column

158

a. Monitoring locations 1, 2, and 3

b. Monitoring locations 4, 5, 6, 7, and 8


Figure 2-79 Reflected overpressure histories; Model VII; spherical charge of 22.68 kg of
TNT; JWL EOS

159

c. Monitoring locations 9, 10 and 11

d. Monitoring locations 12, 13, 14, 15 and 16


Figure 2-79 Reflected overpressure histories; Model VII; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

160

e. Monitoring locations 17, 18 and 19

f. Monitoring locations 20, 21, 22, 23 and 24


Figure 2-79 Reflected overpressure histories; Model VII; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

161

g. Monitoring locations 25, 26 and 27

h. Monitoring locations 28, 29, 30, 31, and 32


Figure 2-79 Reflected overpressure histories; Model VII; spherical charge of 22.68 kg of
TNT; JWL EOS (cont.)

162

a. Net loading

b. Impulse history
Figure 2-80 Net load and impulse histories; Model VII; spherical charge of 22.68 kg of
TNT; JWL EOS

163

2.7.4 Influence of Target Shape on Air-Blast Loadings


Figures 2-81, 2-82, 2-83 and 2-84 enable a comparison of pressures on the front faces of the two
columns at 0, 149, 298 and 447 mm, respectively, above the ground surface, and at y = 0, 160,
320 and 480 mm, respectively. As expected, the overpressure histories at identical locations on
the centerline of the two columns are very similar, as shown in Figures 2-81a, 2-82a, 2-83a and
2-84a. Panels c through d of Figures 2-81 through 2-84 show that the column shape has a
significant effect on the reflected peak overpressure, with large reductions of the circular shape
for a given height and y dimension, due in part to the increased scaled range and increased angle
of incidence.

a. Monitoring locations 1 (Model VI) and 1 (Model VII)


Figure 2-81 Reflected overpressure histories at 0 mm above the ground surface; Models VI
and VII; spherical charge of 22.68 kg of TNT; JWL EOS

164

b. Monitoring locations 2 (Model VI) and 2 (Model VII)

c. Monitoring locations 3 (Model VI) and 3 (Model VII)


Figure 2-81 Reflected overpressure histories at 0 mm above the ground surface; Models VI
and VII; spherical charge of 22.68 kg of TNT; JWL EOS (cont.)

165

d. Monitoring locations 4 (Model VI) and 4 (Model VII)


Figure 2-81 Reflected overpressure histories at 0 mm above the ground surface; Models VI
and VII; spherical charge of 22.68 kg of TNT; JWL EOS (cont.)

166

a. Monitoring locations 11 (Model VI) and 9 (Model VII)

b. Monitoring locations 12 (Model VI) and 10 (Model VII)


Figure 2-82 Reflected overpressure histories at 149 mm above the ground surface;
Models VI and VII; spherical charge of 22.68 kg of TNT; JWL EOS

167

c. Monitoring locations 13 (Model VI) and 11 (Model VII)

d. Monitoring locations 14 (Model VI) and 12 (Model VII)


Figure 2-82 Reflected overpressure histories at 149 mm above the ground surface;
Models VI and VII; spherical charge of 22.68 kg of TNT; JWL EOS (cont.)

168

a. Monitoring locations 21 (Model VI) and 17 (Model VII)

b. Monitoring locations 22 (Model VI) and 18 (Model VII)


Figure 2-83 Reflected overpressure histories at 298 mm above the ground surface;
Models VI and VII; spherical charge of 22.68 kg of TNT; JWL EOS

169

c. Monitoring locations 23 (Model VI) and 19 (Model VII)

d. Monitoring locations 24 (Model VI) and 20 (Model VII)


Figure 2-83 Reflected overpressure histories at 298 mm above the ground surface;
Models VI and VII; spherical charge of 22.68 kg of TNT; JWL EOS (cont.)

170

a. Monitoring locations 31 (Model VI) and 25 (Model VII)

b. Monitoring locations 32 (Model VI) and 26 (Model VII)


Figure 2-84 Reflected overpressure histories at 447 mm above the ground surface;
Models VI and VII; spherical charge of 22.68 kg of TNT; JWL EOS

171

c. Monitoring locations 33 (Model VI) and 27 (Model VII)

d. Monitoring locations 34 (Model VI) and 28 (Model VII)


Figure 2-84 Reflected overpressure histories at 447 mm above the ground surface;
Models VI and VII; spherical charge of 22.68 kg of TNT; JWL EOS (cont.)

172

SECTION 3
INCIDENT AND NORMALLY REFLECTED OVERPRESSURE AND
IMPULSE FOR DETONATIONS IN FREE AIR
3.1 Blast-Resistant Design
Typical blast-resistant design practice, in the United States, United Kingdom, Australia, Canada,
and elsewhere involves the use of empirical air-blast relationships developed by Kingery and his
co-workers (Kingery and Pannill 1964, Kingery 1966, Kingery and Bulmash 1984). These
relationships are presented as charts for spherical free-air and hemispherical surface bursts of
TNT. The Kingery and Bulmash (KB) charts4 are based on high order polynomials (Kingery and
Bulmash 1984) that express incident and normally reflected overpressures, incident and normally
reflected impulses, arrival time and other blast wave parameters as a function of scaled distance,
Z.

3.2 Air-Blast Parameter Studies


3.2.1 Kingery and Pannill, 1964: Hemispherical TNT Surface Bursts
Kingery and Pannill (1964) compiled and processed incident overpressure measurements from
surface detonations of hemispheres of TNT with weights of 5, 20 and 100 tons. The tests were
performed at the Suffield Experimental Station (SES), Alberta Canada. The 5, 20 and 100 ton
shots were executed in 1959, 1960, and 1961, respectively.
Kingery and Pannill noted that the raw data (which is not readily available) was reported in
Groves (1961), a 1961 SES internal report, Kingery et al. (1962), James (1962) and internal
correspondence between Kingery and his SES colleagues. Information on the instrumentation
used is not described by these authors beyond overpressure versus time gages and [a] photooptical shock front velocity technique The gages and the photo-optical technique enable a
calculation of the shock front arrival time and the shock front velocity. The calculated velocity
4

The Kingery and Bulmash charts are reproduced in Appendix D in SI and US units. The air-blast
parameters identified in the legend are defined in Section 1.4.

173

was input to the Rankine-Hugoniot equations to estimate incident peak overpressure. It is not
clear if any direct measurements were made of incident peak overpressure. The authors do not
report the range of scaled distance over which shock front velocities were measured.
Kingery and Pannill developed an equation for the relationship between incident peak
overpressure and scaled distance. Two hundred and seventy three data (likely indirect
measurements of peak overpressure) were used to develop an eighth order polynomial: forty five
data from the 5-ton shot, one hundred and forty data from the 20-ton shot, and eighty eight data
from the 100-ton shot. The authors performed a rudimentary statistical analysis of the proposed
polynomial to characterize its ability to recover the inferred values of incident peak overpressure.
The polynomial is provided for Z 0.5 ft/lb1/3 (0.198 m/kg1/3) but no information is available on
the accuracy of calculation at the low end of the range where the shock front velocity changes
rapidly.
3.2.2 Kingery, 1966: Hemispherical TNT Surface Bursts
In 1966, Kingery extended the presentation of Kingery and Pannill (1964) to more complete
describe blast wave parameters, including shock front arrival time, incident peak overpressure
and impulse, and positive phase duration. Data from the three TNT hemispherical surface bursts
discussed in Kingery and Pannill (1964) and from a 500-ton hemispherical surface-burst shot at
SES in 1964 were evaluated.
Kingery reported the size of the blocks of TNT used to build the hemispherical 500-ton weapon
(305305102 mm); the packing density of the TNT (1560 kg/m3), the soil at the site (glacial
silt with gravel, sand and clay at depth) and the methods used to calculate values of the blast
wave parameters. These methods are summarized below to a) support the presentation of Section
3.2.1, and b) provide insight into the possible uncertainty and variability in the inferred values of
the blast wave parameters.
Shock front arrival times were measured in the 500 ton shot by photo-optical techniques, blast
switches, overpressure gages and slifer (coaxial) cables. The photo-optical technique, which
involves high-speed cameras and angled black-and-white striped (zebra) boards, takes advantage
174

of the differences in density, overpressure and temperature across a shock front. Tracking the
arrival of the shock front at points along the zebra boards enabled a calculation of shock front
velocity and incident peak overpressure. Blast switches produce an electric signal when struck by
a shock front and enable the direct measurement of shock front arrival. Information on the
overpressure gages is not provided and so their operating range and accuracy is unknown.
Seventy-eight data are reported for arrival time in the range 0.208 Z 316 ft/lb1/3 (0.083 Z
126 m/kg1/3).
Kingery noted that incident peak overpressure was recorded directly by pressure transducers and
inferred from measurements of shock front arrival. The data from the 500 ton shot was generally
supportive of the polynomial described in Kingery and Pannill (1964) except for Z < 1.8 ft/lb1/3
(Z < 0.71 m/kg1/3). Positive phase duration was described as difficult to measure because the
pressure transducers were prone to malfunction after the arrival of the shock front. The decay in
some of the pressure histories was extrapolated to ambient. One hundred and thirty-four values
of positive phase duration were reported in the range 0.49 Z 428 ft/lb1/3 (0.194 Z 170
m/kg1/3); only two points were provided for Z < 1.0 ft/lb1/3 (Z < 0.397 m/kg1/3).
Ninety-four values of incident impulse were computed using pressure histories and information
on incident peak overpressure, positive phase duration and rate of decay behind the shock front,
over the range 0.49 Z 428 ft/lb1/3 (0.194 Z 170 m/kg1/3). Only four data points were
provided for Z < 3.0 ft/lb1/3 (Z < 1.19 m/kg1/3).
Hand drawn curves were generated as a function of scaled distance to best fit the available data
for arrival time, incident peak overpressure and impulse, and positive phase duration. A
rudimentary statistical analysis was performed to characterize the relative errors associated with
the use of the curves, noting that the accuracy of the underlying data is unknown.
3.2.3 Kingery and Bulmash, 1984: Spherical Free Air and Hemispherical Surface Bursts of
TNT
Kingery and Bulmash (1984) compiled, analyzed and documented airblast parameters for
spherical free-air bursts and hemispherical surface bursts of TNT in support of a revision to the
175

tri-service manual TM 5-1300 (Department of the Army, the Navy, and the Air Force 1969).
They developed polynomials to express the airblast parameters of arrival time, shock front
velocity, incident peak overpressure and impulse, positive phase duration, and normally reflected
peak overpressure and impulse as a function of scaled distance. The derived polynomials were
reproduced in the form of design charts, which have been published in many government
documents, design standards, and textbooks. The design charts are widely labeled as the
Kingery-Bulmash (KB) charts.
Kingery and Bulmash compiled airblast data from many sources, which although identified in
their Ballistics Research Laboratory report, are generally unavailable and so are not listed here.
The raw data were corrected for charge weight and atmospheric conditions to enable calculations
for a kilogram mass of TNT at atmospheric pressure at sea level (= 101.3 kPa, 14.7 psi).
Kingery and Bulmash described the methods used to measure airblast parameters and this
provides insight into the likely uncertainties in the raw data used to generate the KB polynomials
and charts. Arrival time and incident peak overpressure were measured using the methods
described in Section 3.2.2. Impulse was determined by integrating overpressure over the duration
of the positive phase. The duration of the positive phase was determined from overpressure
histories (and described as one of the more difficult blast parameters to measure with
consistency and repeatability). Reflected overpressure was calculated a) indirectly using theory,
measurements of incident overpressure, and the specific heat ratio, which is affected by
overpressure and temperature, or b) directly from ground measurements of the effects of free-air
bursts. Little information is provided on methods used to calculate reflected impulse although the
authors note there is a lack of specific measurements of this blast parameter.
3.2.3.1 Polynomial Data for TNT Spherical Free-Air Bursts
The figures presented in Kingery and Bulmash, together with an understanding of how data were
collected and processed, enable an assessment of the likely accuracy of the KB polynomials over
the scaled range considered here. Considerable data are provided for incident peak overpressure
but the scatter at small values of scaled distance is significant, which is not surprising given that
the data were inferred and not measured directly. Data for incident peak overpressure are
176

reported in the range 0.134 ft/lb1/3 Z 100 ft/lb1/3 (0.0531 m/kg1/3 Z 40 m/kg1/3). Much less
data were available for incident impulse, with only 3 data points below Z = 0.3 ft/lb1/3 (0.119
m/kg1/3). Kingery and Bulmash noted At scaled distances less than 1 [ft/lb1/3] the measured
values of incident impulse are very few and sometimes of suspect quality. Values of normally
reflected peak overpressure were typically calculated using theory and values of incident peak
overpressure, as noted previously. There were no reflected overpressure data below Z = 0.25
ft/lb1/3 (0.1 m/kg1/3). There were little data for reflected impulse.
3.2.3.2 Polynomial Data for TNT Hemispherical Surface Bursts
Kingery and Bulmash used the incident airblast data of Kingery (1966) to generate polynomials
and a design chart for hemispherical surface bursts of TNT. Values of reflected peak
overpressure were calculated using theory and values of incident peak overpressure. Reflected
impulse was calculated using the spherical free-air burst data and a factor on charge weight of
1.8.
3.2.4 Implementation in UFC 3-340-02 using Design Charts
The KB charts are presented in UFC 3-340-02 (DoD 2008) and are widely used for the protective
design of structures. The incident and normally peak reflected overpressures and impulses
obtained from the KB charts are combined with reflection coefficients that vary as a function of
the angle of incidence to describe the pressure loading history on a surface. Reflection
coefficients are discussed in Chapter 5. Swisdak (1994) and Bogosian et al. (2002) evaluated the
utility of the KB charts, and those studies are described next.

3.3 Studies by Swisdak and Bogosian et al.


Swisdak (1994) reviewed Kingery (1966) and Kingery and Bulmash (1984) for hemispherical
surface bursts of TNT for the purpose of developing alternate equations that would 1) be simpler
to apply than those presented in Kingery and Bulmash (1984), and 2) recover the values of the
KB airblast parameters within 1%. The scope of the Swisdak study did not include a reevaluation of the data used by Kingery and Bulmash to establish their polynomials. Swisdak
provided polynomials for arrival time, incident peak overpressure, normally reflected peak
177

overpressure, positive phase duration, incident impulse, reflected impulse, and shock front
velocity.
Bogosian et al. (2002) sought to 1) compare results from widely used airblast codes (CONWEP,
SHOCK and BlastX) with KB predictions, 2) compare KB predictions with experimental data,
and 3) quantify the uncertainty in test data and make recommendations regarding the use of the
KB charts/predictions. The CONWEP predictions recovered the KB curves for incident and
reflected peak overpressure and impulse exactly, which is expected because CONWEP
implements the KB polynomials. The experimental data assembled by Bogosian et al., across
eleven test series, represented a significant expansion of the dataset available to prior researchers.
The data set comprised a wide range of charge weights, charge geometries, types of explosives,
and heights of burst. Three hundred and three gage records were assembled and analyzed.
Values of scaled distance ranged from approximately 2.5 ft/lb1/3 to 60 ft/lb1/3 (1.0 m/kg1/3 to 24
m/kg1/3) for incident peak overpressure and impulse and from 3.5 ft/lb1/3 to 40 ft/lb1/3 (1.4 m/kg1/3
to 16 m/kg1/3) for reflected peak overpressure and impulse. For incident peak overpressure, the
KB curve recovered the experimental observations well for Z > 10 ft/lb1/3 (4.0 m/kg1/3) but
substantially underpredicted the recorded data for Z < 6 ft/lb1/3 (2.4 m/kg1/3). For incident
impulse, the KB curve overpredicted the recorded data by 15% to 20% across the range 3.0 Z
60 ft/lb1/3 (1.2 Z 24 m/kg1/3). For reflected peak overpressure, the KB curve recovered the
mean of the experimental data reasonably well. For reflected impulse, the KB curve recovered
the measured results well for Z < 7 ft/lb1/3 (2.8 m/kg1/3) but overpredicted the measured results by
a wide margin for Z > 10 ft/lb1/3 (4 m/kg1/3). Bogosian attributed the overprediction of reflected
impulse by the KB curves for Z > 7 ft/lb1/3 (2.8 m/kg1/3) to clearing.

3.4 Evaluation of the KB Charts


The pressure inside an ideal high explosive after detonation is defined by the Chapman-Jouguet
(CJ) point, which describes the minimum propagation velocity of the detonation front for the
reacting gases to reach sonic velocity within the material. The CJ detonation pressure is
associated with this condition, which is 21 GPa for the detonation of TNT of normal packing
density (1630 kg/m) in air at ambient pressure (Baker et al. 1983).

178

At the face of a spherical TNT charge (Z = 0.0527 m/kg1/3) in free air, the KB chart predicts an
incident peak overpressure of 50 MPa for a spherical free-air burst. This overpressure is a very
small fraction (= 1/420) of the CJ pressure of 21 GPa, suggesting a discontinuity in the flow field,
which is physically impossible. The lack of direct measurements of overpressure, impulse,
arrival time and positive phase duration in the near field, as described in prior sections of this
chapter, and the significant discrepancy at the face of the charge noted above, calls into question
the accuracy of the KB charts, especially in the near field.
The remaining sections of this chapter examine incident and normally reflected overpressures
and impulses and shock-front arrival time for spherical free-air bursts of TNT. The near field is
emphasized because many of the threats considered nowadays for security design are associated
with small standoff distances. The near field is (loosely) defined here as the zone within
approximately 7 charge radii ( r 7, Z < 0.4 m/kg1/3) of the point of detonation, where the blast
wave has not yet fully formed and the expansion of the products of combustion has a significant
effect on the blast wave parameters. In this region, direct measurement of pressure and impulse
using commercially available transducers is not possible for the reasons given in Section 2.2.
Analysis is performed using the verified and validated computational fluid dynamics (CFD) code
AUTODYN

(ANSYS 2013a) to 1) determine incident and normally reflected peak

overpressures and impulses, and arrival time of the shock front, and to 2) judge the accuracy of
the KB charts.

3.5 Numerical Studies


3.5.1 Spherical Charges used for Numerical Analysis
A two-phase fluid is modeled numerically, comprising the explosive (combustion products) and
the surrounding air. The ideal high explosive TNT is used for the CFD computations. Air is
modeled as an ideal gas. Radial symmetry is used for analysis of the early expansion of the
combustion products. The Jones-Wilkins-Lee (JWL) Equation of State (EOS) is employed for
this purpose using the properties described in Dobratz and Crawford (1985); see Table 2-1. The
5

AUTODYN was verified and validated in Section 2.3.

179

chemistry of the explosion is not modeled. Afterburning is not modeled because afterburning
energy may not be realized for explosions of unconfined TNT due to incomplete mixing of fuel
with available oxygen, and temperature less than the ignition temperature of the TNT fuels (
1800 K) (McNesby et al. 2010).
Three TNT charge weights (18735 kg, 960 kg and 23 kg) are used for the simulations. The TNT
charge weight of 18735 kg is based on the charge radius of 1.4 m used for the Needham analysis,
presented in Chapter 2, and a TNT packing density of 1630 kg/m3. Twenty-three kilograms is
equally to approximately 50 pounds, which is a typical satchel-type weapon considered in
security design. The 960 kg (2116 pounds) charge could be representative of a vehicle-borne
weapon. The packing density for the 960 and 23 kg charges is 1630 kg/m3, and the
corresponding charge radii are 0.520 m and 0.150 m, respectively. Simulations are performed
with three charge weights to determine the effect, if any, of charge weight on the normalized
blast wave parameters.
Preliminary analysis and a review of the literature (e.g., Needham 2010) showed that
overpressure, particle velocity and particle acceleration change very rapidly close to the face of a
charge. The following section provides an assessment of overpressures and hydrodynamic
parameters close to the face of a charge and establishes a minimum value of scaled distance for
the subsequent CFD analyses.
3.5.2 Overpressures, Particle Velocities and Accelerations Near the Charge Face
Incident peak overpressure changes rapidly in close proximity to the face of a charge, as seen in
Chapter 2 and Appendix C. To explore these changes, and to help frame the presentations later in
this chapter and in Chapter 4, incident overpressures, particle velocities and particle accelerations
are calculated for two cell sizes (0.05 mm and 0.1 mm) and two charge weights (23 kg and 960
kg). Results are plotted as a function of scaled distance in Figures 3-1, 3-2, 3-3, 3-4, 3-5 and 3-6
to examine changes in the immediate vicinity of the face of the charge, namely, 0.9 r 1.1
(0.0474 Z 0.0580 m/kg1/3), where r = 1 at Z = 0.0527 m/kg1/3. Finer meshes are needed to
resolve particle velocity and acceleration than for overpressure. Particle velocity is a function of
differential pressure. Particle acceleration is derived from particle velocity. The calculated
180

particle acceleration and particle velocity very close to the face of the charge are dependent on
cell size and change rapidly. Values stabilize at Z = 0.0553 m/kg1/3 ( r = 1.05) and are similar for
the two charge weights. Peak particle accelerations at the face of the charge, where the pressure
differential is greatest, are on the order of 1011 g ( 1012 m/s2), which was noted in Section 2.2
and is consistent with the range reported by Needham (2010).
A minimum value of Z = 0.0553 m/kg1/3 ( r = 1.05) is used for the CFD calculations presented in
the remaining sections of this chapter and in Chapter 4. The effect on design practice of not
reporting pressures and impulses closer to the face of the charge than a radial expansion of 1.05
is insignificant, as discussed in Section 4.5.

181

Linear scale

Log scale

(a) Overpressure

(b) Particle velocity

(c) Particle acceleration


Figure 3-1 Incident overpressures and particle velocities and accelerations near the face of
the charge ; TNT weight of 23 kg; 0.1 mm cells

182

Linear scale

Log scale

(a) Overpressure

(b) Particle velocity

(c) Particle acceleration


Figure 3-2 Incident overpressures and particle velocities and accelerations near the face of
the charge ; TNT weight of 23 kg; 0.05 mm cells

183

Linear scale

Log scale

(a) Overpressure

(b) Particle velocity

(c) Particle acceleration


Figure 3-3 Incident overpressures and particle velocities and accelerations near the face of
the charge ; TNT weight of 23 kg; 0.1 and 0.05 mm cells

184

Linear scale

Log scale

(a) Overpressure

(b) Particle velocity

(c) Particle acceleration


Figure 3-4 Incident overpressures and particle velocities and accelerations near the face of
the charge; TNT weight of 960 kg; 0.1 mm cells

185

Linear scale

Log scale

(a) Overpressure

(b) Particle velocity

(c) Particle acceleration


Figure 3-5 Incident overpressures and particle velocities and accelerations near the face of
the charge; TNT weight of 960 kg; 0.05 mm cells

186

Linear scale

Log scale

(a) Overpressure

(b) Particle velocity

(c) Particle acceleration


Figure 3-6 Incident overpressures and particle velocities and accelerations near the face of
the charge; TNT weight of 960 kg; 0.1 and 0.05 mm cells

187

3.5.3 Mesh Sensitivity Analysis


Mesh sensitivity analysis for the simulations of incident and normally reflected overpressures
and impulses is performed using h-refinement (Cook 2002) to a) determine a reasonable balance
between solution accuracy and computational effort, and b) provide guidance to an analyst on the
required cell size for CFD analysis of detonations. Mesh convergence is assumed when the
results obtained using a cell size reduced by a factor of two change by less than 10%.
Simulations of incident overpressure and impulse are performed in a 1D domain using axial
symmetry and the multi-material Euler-Godunov solver (Godunov 1959). The Euler-FCT solver
(e.g., Boris and Book 1971) is not available in AUTODYN for 1D calculations. The fluid is
modeled as described in Section 3.5.1. When the products of combustion have expanded to
approximately 10 times their original volume, the JWL EOS is replaced by the ideal gas EOS to
prevent possible numerical errors, as discussed in Section 2.2.1.
Table 3-1 summarizes the analyses that comprise the AUTODYN dataset for incident peak
overpressure, impulse, and arrival time, subdivided into twenty-six intervals of scaled distance
between 0.0553 Z 40.0 m/kg1/3, where the basis for lower limit on Z of 0.0553 m/kg1/3 is
provided in Section 3.5.2. The intervals correspond to the log-scaled distances seen in the KB
charts that are reproduced in UFC 3-340-02. Each interval listed in Table 3-1 has data at no less
than ten monitoring locations (different values of scaled distance) to enable accurate
interpretation of the numerical data over the full range of scaled distance considered here.

188

Table 3-1 AUTODYN dataset for incident calculations


1/3
Z i Z Z j (m/kg )

Number of monitoring locations

0.0553 Z < 0.060

10

0.060 Z < 0.070

20

0.070 Z < 0.080

20

0.080 Z < 0.090

20

0.090 Z < 0.10

20

0.10 Z < 0.20

20

0.20 Z < 0.30

20

0.30 Z < 0.40

20

0.40 Z < 0.50

10

0.50 Z < 0.60

10

0.60 Z < 0.70

10

0.70 Z < 0.80

10

0.80 Z < 0.90

10

0.90 Z < 1.0

10

1.0 Z < 2.0

10

2.0 Z < 3.0

10

3.0 Z < 4.0

10

4.0 Z < 5.0

10

5.0 Z < 6.0

10

6.0 Z < 7.0

10

7.0 Z < 8.0

10

8.0 Z < 9.0

10

9.0 Z < 10.0

10

10.0 Z < 20.0

10

20.0 Z < 30.0

10

30.0 Z < 40.0

10

189

The 960 kg charge is used for the mesh sensitivity study of incident peak overpressure and
impulse. The influence of charge size on the results is investigated at selected scaled distances
using 18735 and 23 kg charges. Tables 3-2, 3-3 and 3-4 present the cell sizes in each interval
used for analysis of incident peak overpressures and impulses, for charge weights of 960, 18735
and 23 kg, respectively. Each interval is defined by bounds on the scaled distance, Z. The upper
bound on Z in a given interval is transformed to a distance in mm, R j . Four cell sizes are
selected for each interval. Three meshes with cell sizes smaller than R j /500 are analyzed
because a cell size of R j /500 is expected to provide converged solutions based on the studies
presented in Section 2.2. One mesh with a cell size equal to or greater than R j /500 is analyzed
for the first twenty-four intervals because R j /500 does not enable exact modeling of the
geometry (radius) of the charge. The smallest cell size in last two intervals, 20.0 Z < 40.0
m/kg1/3, is less than R j /500 because the distance R j /500 exceeds the charge radius and the
charges must be modeled explicitly.
Simulations of normally reflected peak overpressure and impulse are performed in a twodimensional (2D) domain using the Euler-FCT solver, which is preferred to the Euler-Godunov
solver for reasons of accuracy (Borve et al. 2008). The results from 1D analysis are remapped
into a one-quarter of the 2D air domain, as shown in Figure 3-7, to reduce the computational
expense of the reflected peak overpressure and impulse calculations. For the reflected peak
overpressure and impulse calculations reported here, the 1D results calculated using the smallest
cell size in the interval considered are used as input to the 2D domain. The vertical dimension of
the 2D plane is set equal to the horizontal dimension to avoid spurious reflections from the upper
transmitting boundary as noted in Section 2.5. Normally reflected peak overpressure and impulse
are calculated at monitoring location 1, as identified in Figure 3-7.

190

Table 3-2 Cell sizes for incident peak overpressure and impulse; 960 kg charge with a
radius of 520 mm
1/3
Z i Z Z j (m/kg )

R j Z j W 1/3 (mm)

R j /500 (mm)

Cell size (mm)

0.0553 Z < 0.060

592

1.18

1.18, 1, 0.5 and 0.25

0.060 Z < 0.070

691

1.38

1.38, 1, 0.5 and 0.25

0.070 Z < 0.080

789

1.58

1.58, 1, 0.5 and 0.25

0.080 Z < 0.090

888

1.78

1.78, 1, 0.5 and 0.25

0.090 Z < 0.10

986

1.97

1.98, 1, 0.5 and 0.25

0.10 Z < 0.15

1480

2.96

2.97, 1, 0.5 and 0.25

0.15 Z < 0.20

1973

3.95

3.97, 1, 0.5 and 0.25

0.20 Z < 0.25

2466

4.93

4.95, 1, 0.5 and 0.25

0.25 Z < 0.30

2959

5.92

5.98, 1, 0.5 and 0.25

0.30 Z < 0.40

3946

7.89

8.00, 1, 0.5 and 0.25

0.40 Z < 0.50

4932

9.86

10.00, 8, 4 and 2

0.50 Z < 0.60

5919

11.8

12.1, 8, 4 and 2

0.60 Z < 0.70

6905

13.8

14.1, 8, 4 and 2

0.70 Z < 0.80

7892

15.8

16.3, 8, 4 and 2

0.80 Z < 0.90

8878

17.8

17.9, 15.8, 8 and 4

0.90 Z < 1.0

9865

19.7

20.0, 15.8, 8 and 4

1.0 Z < 1.2

11838

23.7

24.8, 15.8, 8 and 4

1.2 Z < 1.5

14797

29.6

30.6, 15.8, 8 and 4

1.5 Z < 2.0

19730

39.5

40.0, 15.8, 8 and 4

2.0 Z < 3.0

29595

59.2

65.0, 40, 20 and 10

3.0 Z < 4.0

39459

78.9

86.7, 40, 20 and 10

4.0 Z < 5.0

49324

98.6

104, 65, 32.5 and 16.25

5.0 Z < 6.0

59189

118

130, 65, 32.5 and 16.25

6.0 Z < 7.0

69054

138

173, 65, 32.5 and 16.25

7.0 Z < 8.0

78919

158

173, 65, 32.5 and 16.25

8.0 Z < 9.0

88784

178

260, 65, 32.5 and 16.25

9.0 Z < 10.0

98648

197

260, 65, 32.5 and 16.25

10.0 Z < 20.0

197297

395

520, 65, 32.5 and 16.25

20.0 Z < 30.0

295945

592

520, 65, 32.5 and 16.25

30.0 Z < 40.0

394594

789

520, 65, 32.5 and 16.25

191

Table 3-3 Cell sizes for incident peak overpressure and impulse; 18735 kg charge with a
radius of 1400 mm
1/3
Z i Z Z j (m/kg )

R j Z j W 1/3 (mm)

R j /500 (mm)

Cell size (mm)

0.0553 Z < 0.060

1594

3.19

3.19, 2, 1 and 0.5

0.060 Z < 0.070

1859

3.72

3.72, 2, 1 and 0.5

0.070 Z < 0.080

2125

4.25

4.26, 2, 1 and 0.5

0.080 Z < 0.090

2390

4.78

4.79, 2, 1 and 0.5

0.090 Z < 0.10

2656

5.31

5.32, 2, 1 and 0.5

0.10 Z < 0.15

3984

8.00

8.00, 2, 1 and 0.5

0.15 Z < 0.20

5312

10.6

10.7, 2, 1 and 0.5

0.20 Z < 0.25

6640

13.3

13.3, 2, 1 and 0.5

0.25 Z < 0.30

7968

15.9

16.1, 2, 1 and 0.5

0.30 Z < 0.40

10624

21.2

21.5, 2, 1 and 0.5

0.40 Z < 0.50

13280

26.6

26.9, 20, 10 and 5

1.0 Z < 1.2

31871

63.7

66.7, 50, 25 and 12.5

1.2 Z < 1.5

39839

79.9

82.4, 50, 25 and 12.5

1.5 Z < 2.0

53119

106

108, 50, 25 and 12.5

5.0 Z < 6.0

159356

319

350, 200, 100 and 50

6.0 Z < 7.0

185916

372

467, 200, 100 and 50

7.0 Z < 8.0

212475

425

467, 200, 100 and 50

8.0 Z < 9.0

239034

478

700, 200, 100 and 50

9.0 Z < 10.0

265594

531

700, 200, 100 and 50

10.0 Z < 20.0

531188

1062

1400, 200, 100 and 50

20.0 Z < 30.0

796781

1594

1400, 200, 100 and 50

30.0 Z < 40.0

1062375

2125

1400, 200, 100 and 50

192

Table 3-4 Cell sizes for incident peak overpressure and impulse; 23 kg charge with a radius
of 150 mm
1/3
Z i Z Z j (m/kg )

Rj Z j W 1/3 (mm)

R j /500 (mm)

Cell size (mm)

0.0553 Z < 0.060

171

0.34

0.34, 0.3, 0.15 and 0.075

0.060 Z < 0.070

199

0.40

0.40, 0.3, 0.15 and 0.075

0.070 Z < 0.080

228

0.46

0.46, 0.3, 0.15 and 0.075

0.080 Z < 0.090

256

0.51

0.51, 0.3, 0.15 and 0.075

0.090 Z < 0.10

284

0.57

0.57, 0.3, 0.15 and 0.075

0.10 Z < 0.15

427

0.85

0.86, 0.3, 0.15 and 0.075

0.15 Z < 0.20

569

1.14

1.15, 0.3, 0.15 and 0.075

0.20 Z < 0.25

711

1.42

1.43, 0.3, 0.15 and 0.075

0.25 Z < 0.30

853

1.71

1.72, 0.3, 0.15 and 0.075

0.30 Z < 0.40

1138

2.28

2.31, 0.3, 0.15 and 0.075

0.40 Z < 0.50

1422

2.84

2.88, 2, 1 and 0.5

1.0 Z < 1.2

3413

6.83

7.14, 5, 2.5 and 1.25

1.2 Z < 1.5

4266

8.53

8.82, 5, 2.5 and 1.25

1.5 Z < 2.0

5688

11.4

11.5, 5, 2.5 and 1.25

5.0 Z < 6.0

17063

34.1

37.5, 25, 12.5 and 6.25

6.0 Z < 7.0

19907

39.8

50, 25, 12.5 and 6.25

7.0 Z < 8.0

22751

45.5

50, 25, 12.5 and 6.25

8.0 Z < 9.0

25595

51.2

75, 25, 12.5 and 6.25

9.0 Z < 10.0

28439

56.9

75, 25, 12.5 and 6.25

10.0 Z < 20.0

56877

114

150, 25, 12.5 and 6.25

20.0 Z < 30.0

85316

171

150, 25, 12.5 and 6.25

30.0 Z < 40.0

113755

228

150, 25, 12.5 and 6.25

193

Figure 3-7 AUTODYN 2D simulation using 1D remapped data for Z = 1.0 m/kg1/3
Fewer data are calculated for normally reflected peak overpressure and impulse, as shown in
Table 3-5, because a) one 2D simulation is required to calculate reflected peak overpressure and
reflected impulse at a monitoring location (e.g., point 1 in Figure 3-7) whereas a family of
incident peak overpressures and impulses can be derived from a single 1D analysis, and b)
evaluation of the results of the incident peak overpressure and impulse analysis indicated that
fewer simulations are needed to develop robust conclusions. For Z < 0.4 m/kg1/3, calculations are
performed at no less than five values of scaled distance in each interval. For Z 0.4 m/kg1/3,
simulations are performed at the boundaries of the intervals of scaled distance identified in Table
3-1.
Tables 3-6, 3-7 and 3-8 present the cell sizes used to calculate reflected data for the 960, 18735
and 23 kg charges, respectively. The results of the 1D analysis for the smallest cell size are used
as input to the 2D analysis. (For example, the smallest cell size used for the 1D analysis of the
194

960 kg charge for 0.0553 Z < 0.060 m/kg1/3 is 0.25 mm; see row 2 of Table 3-2.) Results of this
1D analysis are mapped into the 2D domain for analysis using the four cell sizes (4.36, 2.18, 1.09
and 0.546 mm) shown in row 2 of Table 3-6. Identical to the simulations of incident overpressure
and impulse, most of the simulations are performed with the 960 kg charge. Square cells with
side dimensions of R j /125, R j /250, R j /500 and R j /1000 are used for the 2D calculations.
For Z < 0.4 m/kg1/3 and the 960 kg charge, four cell sizes are considered for analysis at
boundaries of the intervals, and two cell sizes are considered at the monitoring locations within
the interval. Of the four cell sizes at the boundaries of the intervals, the largest three are used to
compute reflected impulse and reflected peak overpressure (e.g., 4.36, 2.18 and 1.09 mm for Z =
0.0553 m/kg1/3) and the smallest (e.g., 0.546 mm) is used to compute reflected peak overpressure
only. This choice of cell sizes is driven by the observation that the largest three cell sizes in each
interval produce nearly identical values of reflected impulse, whereas only the smallest two cell
sizes produced values of reflected peak overpressure within 10% of the other, with the smallest
cell size assumed to produce the true (or correct) result. In the intervals for Z < 0.4 m/kg1/3,
reflected impulse is calculated for the larger of the two cell sizes (taken from the converged
solution at the lower limit of the interval) and reflected peak overpressure is calculated for the
smaller of the two cell sizes (taken from the converged solution at the lower limit of the interval).
Consider Z = 0.0553 m/kg1/3 in Table 3-6. The reflected impulse converged for a cell size of
larger than 2.18 mm and the reflected peak overpressure converged for a cell size greater than
1.09 mm. In the interval with Z = 0.0553 m/kg1/3 as the lower limit, cell sizes of 1.09 and 0.546
mm are used for the impulse and overpressure calculations, respectively, thereby ensuring a
converged solution.
The influence of charge size on the reflected peak overpressure and impulse is investigated at
selected scaled distances using the 18735 and 23 kg charges, as shown in Tables 3-7 and 3-8,
respectively. Identical to the analysis of the 960 kg charge, of the four cell sizes listed at each
scaled distance in Tables 3-7 and 3-8, the largest three ( R j /125, R j /250, R j /500) are used to
compute reflected impulse and reflected peak overpressure and the smallest ( R j /1000) is used to
compute reflected peak overpressure only.

195

Table 3-5 AUTODYN dataset for reflected calculations


Z j Z Zk

or Z Z j (m/kg1/3)

Number of monitoring locations

0.0553 Z < 0.060

0.060 Z < 0.070

0.070 Z < 0.080

0.080 Z < 0.090

0.090 Z < 0.10

0.10 Z < 0.20

0.20 Z < 0.30

0.30 Z < 0.40

Z = 0.40

Z = 0.50

Z = 0.60

Z = 0.70

Z = 0.80

Z = 0.90

Z = 1.0

Z = 2.0

Z = 3.0

Z = 4.0

Z = 5.0

Z = 6.0

Z = 7.0

Z = 8.0

Z = 9.0

Z = 10.0

Z = 20.0

Z = 30.0

Z = 40.0

196

Table 3-6 Cell sizes for reflected peak overpressure and impulse; 960 kg charge with a
radius of 520 mm
Z j Z Zk

or Z Z j (m/kg1/3)

Z = 0.0553
0.0553 < Z < 0.060
Z = 0.060
0.060 Z < 0.070
Z = 0.070
0.070 Z < 0.080
Z = 0.080
0.080 Z < 0.090
Z = 0.090
0.090 Z < 0.10
Z = 0.10
0.10 Z < 0.20
Z = 0.20
0.20 Z < 0.30
Z = 0.030
0.30 Z < 0.40
Z = 0.40
Z = 0.50
Z = 0.60
Z = 0.70
Z = 0.80
Z = 0.90
Z = 1.0
Z = 2.0
Z = 3.0
Z = 4.0
Z = 5.0
Z = 6.0
Z = 7.0
Z = 8.0
Z = 9.0
Z = 10.0
Z = 20.0
Z = 30.0
Z = 40.0

Rj Z j W 1/3 (mm)

R j /500 (mm)

Cell size (mm)

546
546
592
592
691
691
789
789
888
888
986
986
1973
1973
2959
2959
3946
4932
5919
6905
7892
8878
9865
19730
29595
39459
49324
59189
69054
78919
88784
98648
197297
295945
394594

1.09
1.09
1.18
1.18
1.38
1.38
1.58
1.58
1.78
1.78
1.97
1.97
3.95
3.95
5.92
5.92
7.89
9.87
11.8
13.8
15.8
17.8
19.7
39.5
59.2
78.9
98.6
118
138
158
178
197
395
592
789

4.36, 2.18, 1.09 and 0.546


1.09 and 0.546
4.74, 2.37, 1.18 and 0.592
1.18 and 0.592
5.52, 2.76, 1.38 and 0.691
1.38 and 0.691
6.31, 3.16, 1.58 and 0.789
1.58 and 0.789
7.10, 3.55, 1.78 and 0.888
1.78 and 0.888
7.89, 3.95, 1.97 and 0.986
1.97 and 0.986
15.8, 7.89, 3.95 and 1.98
3.95 and 1.98
23.7, 11.8, 5.92 and 2.96
5.92 and 2.96
31.6, 15.8, 7.89 and 3.95
39.46, 19.7, 9.87 and 4.93
47.4, 23.7, 11.8 and 5.92
55.2, 27.6, 13.8 and 6.90
63.1, 31.6, 15.8 and 7.89
71.0, 35.5, 17.8 and 8.88
78.9, 39.5, 19.7 and 9.87
158, 78.9, 39.5 and 19.7
237, 118, 59.2 and 29.6
316, 158, 78.9 and 39.5
395, 197, 98.6 and 49.3
474, 237, 118 and 59.2
552, 276, 138 and 69.0
632, 316, 158 and 78.9
710, 355, 178 and 88.8
789, 395, 197 and 98.7
1578, 789, 395 and 197
2368, 1184, 592 and 296
3157, 1578, 789 and 395

197

Table 3-7 Cell sizes for reflected peak overpressure and impulse; 18735 kg charge with a
radius of 1400 mm
1/3
Z Z j (m/kg )

Rj Z j W 1/3 (mm)

R j /500 (mm)

Cell size (mm)

Z = 0.0553

1469

2.94

11.8, 5.88, 2.94 and 1.47

Z = 0.060

1594

3.19

12.7, 6.37, 3.19 and 1.59

Z = 0.080

2125

4.25

17.0, 8.50, 4.25 and 2.13

Z = 0.10

2656

5.31

21.2, 10.6, 5.31 and 2.66

Z = 0.20

5312

10.6

42.5, 21.2, 10.6 and 5.31

Z = 0.40

10624

21.2

85.0, 42.5, 21.2 and 10.6

Z = 1.0

26559

53.1

212, 106, 53.1 and 26.6

Z = 5.0

131797

266

1062, 531, 266 and 133

Z = 10.0

265594

531

2125, 1062, 531 and 266

Z = 40.0

1062375

2125

8500, 4250, 2125 and 1062

Table 3-8 Cell sizes for reflected peak overpressure and impulse; 23 kg charge with a radius
of 150 mm
1/3
Z Z j (m/kg )

Rj Z j W 1/3 (mm)

R j /500 (mm)

Cell size (mm)

Z = 0.0553

157

0.315

1.26, 0.629, 0.315 and 0.157

Z = 0.060

171

0.341

1.37, 0.683, 0.341 and 0.171

Z = 0.080

228

0.455

1.82, 0.910, 0.455 and 0.228

Z = 0.10

284

0.569

2.28, 1.14, 0.569 and 0.284

Z = 0.20

569

1.14

4.55, 2.28, 1.14 and 0.569

Z = 0.40

1138

2.28

9.10, 4.55, 2.28 and 1.14

Z = 1.0

9865

5.69

22.8, 11.4, 5.69 and 2.84

Z = 5.0

14217

28.4

114, 56.9, 28.4 and 14.2

Z = 10.0

28439

56.9

228, 114, 56.9 and 28.4

Z = 40.0

113755

228

910, 455, 228 and 114

198

3.6 Incident Overpressure and Impulse


3.6.1 Results and Observations
Figures 3-8, 3-9, 3-10, 3-11 and 3-12 present results of calculations for incident peak
overpressure and impulse. Figures 3-8 and 3-9 present the ratios of incident peak overpressure
for the different cell sizes of Tables 3-2, 3-3 and 3-4. The legends in the panels identify three
ratios that involve the R j +/500 (largest, greater than R j /500), coarse (second largest), medium
(third largest) and fine (smallest) meshes. The coarse, medium and fine meshes involve cells
smaller than R j /500. The ratios of incident peak overpressure for the coarse-to-fine and
medium-to-fine meshes are less than 10% at all scaled distances, namely, 0.0553 Z < 40
m/kg1/3, as seen in Figure 3-8. The ratios for the R j /500-to-fine meshes are mostly less than 10%
for Z < 8.0 m/kg1/3. For incident impulse, the three ratios are less than 10% at all scaled distances
as seen in Figure 3-9. Figure 3-10 presents the ratios of incident peak overpressure and scaled
incident impulse for the three charge weights (18735, 960 and 23 kg) and the smallest cells used
for its analysis. The effect of charge weight is clearly insignificant.
The incident peak overpressures calculated for the three charge weights, using the smallest cells
in the intervals of Table 3-2, 3-3 and 3-4, are presented in Figure 3-11. The KB predictions and
the data of Needham (2010) are included in the figures. Figure 3-11a presents data across the
range 0.0553 Z 40 m/kg1/3. Figures 3-11b through 3-11f present data across smaller ranges of
scaled distance to aid interpretation. Figures 3-11g and 3-11h present the ratios of the
AUTODYN-calculated to the KB-predicted values of incident peak overpressure for 0.0553 Z
40 m/kg1/3, where the truncated y axis of Figure 3-11h allows the reader to judge the range over
which the ratio is between 0.9 and 1.1. Figure 3-12 presents information for incident impulse,
which are also calculated using the smallest cell size in the intervals of Tables 3-2, 3-3, and 3-4.

199

(a) 18735 kg charge

(b) 960 kg charge

(c) 23 kg charge
Figure 3-8 Ratios of incident peak overpressure for different cell sizes; 0.0553 Z < 40.0
m/kg1/3
200

(a) 18735 kg charge

(b) 960 kg charge

(c) 23 kg charge
Figure 3-9 Ratios of incident impulse for different cell sizes; 0.0553 Z < 40.0 m/kg1/3

201

(a) Charge weights of 18735 and 960 kg

(b) Charge weights of 960 and 23 kg

(c) Charge weights of 18735 and 23 kg


Figure 3-10 Ratios of incident peak overpressure and impulse for the fine meshes; charge
weights of 18735, 960 and 23 kg; 0.0553 Z < 40.0 m/kg1/3
202

(a) Incident overpressure; 0.0553 Z 40.0 m/kg1/3

(b) Incident overpressure; 0.0553 Z 0.06 m/kg1/3


Figure 3-11 Incident peak overpressure versus scaled distance

203

(c) Incident overpressure; 0.06 Z 0.1 m/kg1/3

(d) Incident overpressure; 0.1 Z 1 m/kg1/3


Figure 3-11 Incident peak overpressure versus scaled distance (cont.)

204

(e) Incident overpressure; 1 Z 10 m/kg1/3

(f) Incident overpressure; 10 Z 40 m/kg1/3


Figure 3-11 Incident peak overpressure versus scaled distance (cont.)

205

(g) Ratios of AUTODYN calculations to KB predictions

(h) Ratios of AUTODYN calculations to KB predictions; part range on ratio


Figure 3-11 Incident peak overpressure versus scaled distance (cont.)
206

(a) Incident impulse; 0.0553 Z 40.0 m/kg1/3

(b) Incident impulse; 0.0553 Z 0.06 m/kg1/3


Figure 3-12 Incident impulse versus scaled distance

207

(c) Incident impulse; 0.06 Z 0.1 m/kg1/3

(d) Incident impulse; 0.1 Z 1 m/kg1/3


Figure 3-12 Incident impulse versus scaled distance (cont.)

208

(e) Incident impulse; 1 Z 10 m/kg1/3

(f) Incident impulse; 10 Z 40 m/kg1/3


Figure 3-12 Incident impulse versus scaled distance (cont.)

209

(g) Ratio of AUTODYN calculations to KB predictions; full range

(h) Ratios of AUTODYN calculations to KB predictions; part range on ratio


Figure 3-12 Incident impulse versus scaled distance (cont.)
210

The AUTODYN calculations of incident peak overpressure are significantly greater than the KB
predictions for Z < 0.08 m/kg1/3 but are within 10% (generally) for Z > 0.08 m/kg1/3. The
maximum difference is a factor of 14.2 at Z = 0.0553 m/kg1/3. For incident impulse, the ratio of
the AUTODYN calculations to the KB predictions is generally either greater than 1.1 or less than
0.9. The maximum difference in incident impulse is a factor of 3.52 at Z = 0.0553 m/kg1/3. In the
far field, say Z > 1.0 m/kg1/3, the KB predictions of incident impulse are 10% to 25% greater than
the AUTODYN calculations, which is consistent with the observations of Bogosian et al. (2002),
(see Section 3.3), and with the analysis of Browning et al. (2013) using the computer codes LSDYNA (LSTC 2013) and CTH (McGlaun et al. 1990), as discussed in Section 2.3.
3.6.2 Cell Size for CFD Analysis
One goal of this analysis was to provide recommendations on the choice of cell size for CFD
analysis of detonation of high explosive.
For incident peak overpressure, a cell size equal to 0.002 times the distance to the monitoring
location will provide results within 10% of the converged value for Z < 8.0 m/kg1/3, noting that
the choice of cell size must allow the adequate meshing of the explosive. For Z > 8.0 m/kg1/3, a
cell size smaller than 0.002 times the distance should be used to provide results within 10% of
the converged value. (At Z = 8.0 m/kg1/3, the incident peak overpressure is approximately oneseventh of an atmosphere and the use of CFD tools to predict incident pressure and impulse is
likely not warranted.) For incident impulse, a cell size equal to 0.002 times the distance to the
monitoring location will provide results within 10% of the converged value across the range
0.0553 Z < 40 m/kg1/3.
In summary, a cell size of 0.002 times the distance from the point of detonation to the monitoring
location is recommended for CFD analysis, noting that CFD analysis is likely not warranted for
blast-resistant design at large values of scaled distance because the incident peak overpressure is
small. The choice of cell size must accommodate the meshing of the charge and this may dictate
the use of a smaller cell size than that required for the accurate calculation of values of peak
overpressure and impulse.

211

3.7 Reflected Overpressure and Impulse


3.7.1 Results and Observations
Figures 3-13, 3-14, 3-15, 3-16 and 3-17 present results of calculations of normally reflected peak
overpressure and impulse. Figure 3-13 presents the ratios of normally reflected peak
overpressure for the coarse-to-very fine, medium-to-very fine, and fine-to-very fine cells. The
ratios for R j /500 (fine) to R j /1000 (very fine) range between 0.9 and 1.1 and so a cell size of

R/500 is considered appropriate for simulating reflected peak overpressure in a 2D domain.


Figure 3-14 presents the corresponding data for reflected impulse using coarse, medium and fine
meshes. The ratios range between 0.95 and 1.05 for the three charge weights and all values of
scaled distance, and so the cell size of R/500 is also considered appropriate for simulating
reflected impulse in a 2D domain.
Figure 3-15 presents ratios of normally reflected peak overpressure and scaled reflected impulse
for the three charge weights (18735, 960 and 23 kg) and the smallest cells used for analysis. The
effect of charge weight is not significant.
The normally reflected peak overpressures for the three charge weights, using the R j /1000 cells,
are presented in Figures 3-16a through 3-16d. The KB predictions are included in the figures.
Figure 3-16a presents data across the range 0.0553 Z 40 m/kg1/3. Figures 3-16b through 316d present data across smaller ranges of scaled distance to aid interpretation. Figure 3-16e and
3-16f present the ratios of the AUTODYN-calculated to the KB-predicted values of normally
reflected peak overpressure for 0.0553 Z 40 m/kg1/3, where the truncated y axis of Figure 316f allows the reader to judge the range over which the ratio is between 0.9 and 1.1. Figure 3-17
presents information for reflected impulse, which are also calculated using the R j /500 cell size
in the intervals of Tables 3-6, 3-7, and 3-8.
The AUTODYN calculations of normally reflected peak overpressure are significantly greater
than the KB predictions for Z < 0.30 m/kg1/3 but within 10% (generally) for Z > 0.30 m/kg1/3.
The maximum difference is a factor of 20 at Z = 0.0553 m/kg1/3. The maximum difference in
reflected impulse is approximately 10% across the entire range of scaled distance.
212

(a) 18735 kg charge

(b) 960 kg charge

(c) 23 kg charge
Figure 3-13 Ratios of reflected peak overpressure for different cell sizes; 0.0553 Z < 40.0
m/kg1/3
213

(a) 18735 kg charge

(b) 960 kg charge

(c) 23 kg charge
Figure 3-14 Ratios of reflected impulse for different cell sizes; 0.0553 Z < 40.0 m/kg1/3

214

(a) Charge weights of 18735 and 960 kg

(b) Charge weights of 960 and 23 kg

(c) Charge weights of 18735 and 23 kg


Figure 3-15 Ratios of reflected peak overpressure and impulse for the smallest cell sizes;
charge weights of 18735, 960 and 23 kg; 0.0553 Z < 40.0 m/kg1/3
215

(a) Reflected overpressure; 0.0553 Z 40.0 m/kg1/3

(b) Reflected overpressure; 0.0553 Z 0.1 m/kg1/3


Figure 3-16 Normally reflected peak overpressure versus scaled distance
216

(c) Reflected overpressure; 0.1 Z 1 m/kg1/3

(d) Reflected overpressure; 1 Z 40 m/kg1/3


Figure 3-16 Normally reflected peak overpressure versus scaled distance (cont.)
217

(e) Ratio of AUTODYN calculations to KB predictions; full range

(f) Ratios of AUTODYN calculations to KB predictions; part range on ratio


Figure 3-16 Normally reflected peak overpressure versus scaled distance (cont.)
218

(a) Reflected impulse; 0.0553 Z 40.0 m/kg1/3

(b) Reflected impulse; 0.0553 Z 0.1 m/kg1/3


Figure 3-17 Normally reflected impulse versus scaled distance
219

(c) Reflected impulse; 0.1 Z 1 m/kg1/3

(d) Reflected impulse; 1 Z 40 m/kg1/3


Figure 3-17 Normally reflected impulse versus scaled distance (cont.)
220

(e) Ratios of AUTODYN calculations to KB predictions


Figure 3-17 Normally reflected impulse versus scaled distance (cont.)
3.7.2 Cell Size for CFD Analysis
It is not straightforward to provide code-independent recommendations for 2D CFD analysis of
the effects of detonations on rigid reflecting surfaces because a) codes employ different
algorithms to map 1D results into a 2D domain, b) the cell size chosen for the 1D analysis will
influence the initial conditions for the 2D analysis, regardless of the cell sizes chosen for the 2D
analysis, c) the choice of perpendicular dimension will affect the accuracy of the results, and d)
3D analysis is generally required to characterize blast loading effects on components of
structures. For 2D analysis using re-mapped 1D results as initial conditions, the computational
time will generally be dominated by the 2D calculations and so as small a cell size as is practical
should be used for the 1D analysis. For the calculations presented here, a 2D cell size equal to
0.002 times the distance to the monitoring location on the reflecting surface provided results
within 10% of the converged value for 0.0553 Z < 40 m/kg1/3.

221

3.8 Shock Front Arrival Time


Arrival time is an important parameter for analysis and design when pressure histories are being
computed across a reflecting surface such as a structural component or a structural system. The
simulations in Section 3.6 for incident overpressure and impulse are mined here for arrival time
data. Figure 3-18 presents the results of the AUTODYN simulations and the KB predictions for
scaled arrival time. The numerical results and the KB predictions are virtually identical for all
charge weights (18735, 960 and 23 kg) across the considered range of scaled distance, 0.0553
Z < 40.0 m/kg1/3. Figure 3-19 presents the ratio of the AUTODYN-calculated to KB-predicted
arrival times. The effect of charge weight is insignificant as shown in Figures 3-19 and 3-20.

Figure 3-18 AUTODYN calculations and KB predictions for specific arrival time

222

Figure 3-19 Ratios of AUTODYN calculations and KB predictions of specific arrival time

Figure 3-20 Effect of charge weight on specific arrival times calculations: 18735, 960 and 23
kg charges

223

SECTION 4
DESIGN CHARTS AND POLYNOMIALS FOR AIR-BLAST
PARAMETERS
4.1 Introduction
The empirical relationships developed by Kingery and Bulmash (1984) form the basis of design
charts that are reproduced in textbooks and US government documents, including UFC 3-340-02
(DoD 2008). These relationships enable air-blast parameters such as incident peak overpressure
and impulse, normally reflected peak overpressure and impulse, arrival time, and positive phase
duration to be empirically derived for spherical free-air and hemispherical surface bursts of TNT.
Pressure-time curves suitable for analysis and design of structural components and systems can
be constructed using charted values of these air-blast parameters and the Friedlander
overpressure history (see Section 1.1).
The accuracy of the Kingery and Bulmash (KB) charts was discussed in Chapter 3. The
AUTODYN data reported in Chapter 3 are used here to generate a new family of polynomials
and charts suitable for blast-resistant design.

4.2 Accuracy of KB and Numerical Predictions of Blast Parameters


4.2.1 Incident Peak Overpressure and Impulse
The KB (or UFC 3-340-02) predictions of incident overpressure for spherical free-air bursts were
compared to AUTODYN CFD calculations in Chapter 3. The AUTODYN calculations of
incident peak overpressure are more than 10% greater than the KB predictions for Z < 0.08
m/kg1/3 (see Figure 3-11h) and are more than a factor of 10 greater than the KB predictions close
to the face of the charge, at Z = 0.0567 m/kg1/3. There is no meaningful difference between the
AUTODYN calculations and the KB predictions of incident peak overpressure in the range 0.08
Z < 40.0 m/kg1/3.

225

For incident impulse, the AUTODYN-calculated values in the range Z < 0.1 m/kg1/3 are more
than 20% greater than the KB predictions (see Figure 3-12g), with the difference being a factor
of 3.5 at Z = 0.0553 m/kg1/3. For 0.1 Z < 40.0 m/kg1/3, the differences are between 10% and
25%, with the KB charts predicting greater values of impulse than AUTODYN. These CFD
predictions in the far field are consistent with the observations of Bogosian et al. (2002), who
noted that the KB charts overpredicted incident impulse for Z > 1.19 m/kg1/3 (Z > 3.0 ft/lb1/3) (see
Section 3.3).
4.2.2 Normally Reflected Peak Overpressure and Impulse
The normally reflected peak overpressures calculated by AUTODYN are more than 10% greater
than the KB predictions for Z < 0.3 m/kg1/3 (see Figure 3-16f). The differences in normally
reflected peak overpressure are small for Z 0.4 m/kg1/3. The AUTODYN calculations and KB
predictions of reflected impulse differ by 10% or less across the range of scaled distance
considered, namely, 0.0553 Z < 40 m/kg1/3.
4.2.3 Shock Front Arrival Time
There are no meaningful differences in the shock front arrival time between the numerical
calculations and the KB predictions across the range of scaled distance considered, namely,
0.0553 Z < 40.0 m/kg1/3.

4.3 Updated Polynomials and Charts for Overpressure, Impulse and Arrival
Time
4.3.1 Charts for Design
Revisions to the spherical free-air burst charts in UFC 3-340-02 are proposed based on the
discussions and AUTODYN simulations presented in Chapters 2 and 3, and in the preceding
sections of this chapter. Polynomials and charts are generated for incident and normally reflected
overpressure and impulse, and for arrival time, for 0.0553 Z 40.0 m/kg1/3.

226

The lower limit on scaled distance of 0.0553 m/kg1/3 is used because the numerical simulations
stabilize at this value of scaled distance, Z, as noted in Section 3.5.2. A value of Z = 0.0553
m/kg1/3 corresponds to a radial expansion of the shock front, r 1.05 . The design implications
of not providing polynomials and charts for 1.00 r 1.05 are insignificant. See Figure 4-1 that
shows a spherical charge and rigid reflecting surface (hatched) to its right. The dashed, dash-dot,
and dash-dot-dot lines represent radial expansions of 1.025, 1.050 and 1.100, respectively. For
design against terrorist threats, a free-air detonation at a radial expansion of less than 1.05 should
probably be treated as a contact detonation because the gap between the charge and reflecting
surface is negligible.

Figure 4-1 Spherical charge and radial expansions, r , of 1.025, 1.050 and 1.100

227

4.3.2 Polynomials
4.3.2.1 Procedure
Polynomial functions of the form shown in Equation 4-1 are used to present air-blast parameters
based on AUTODYN simulations. The functional form of the equations follows that adopted by
Kingery and Bulmash (1984). Polynomials are derived for incident and normally reflected peak
overpressures and impulses, and arrival time 6 . The constants are computed by least squares
fitting:
(4-1)

Y C 0 C1U C N U N

where Y is the common (base 10) logarithm of the blast parameter, U K0 K1 log(Z ) , C and K
are constants, and N is the order of the polynomial.
The AUTODYN data are fitted using high-order polynomials. High-order polynomials are
required to accommodate the rapid changes in pressure and impulse near the face of the charge.
The order of each polynomial is selected such that the ratio of the polynomial calculation to the
AUTODYN calculation lies close 1.0, which is deemed satisfactory for the purpose of design
and better than the resolution enabled by the charts. Two families of polynomials are presented:
1) 0.0553 Z 40.0 m/kg1/3 describing the complete range of scaled distance considered by
Kingery and Bulmash, and 2) 0.0553 Z < 0.5 m/kg1/3 and 0.5 Z 40 m/kg1/3, such that the
KB polynomials may be substituted just in the range Z < 0.5 m/kg1/3 if preferred.
4.3.2.2 Proposed Polynomials for Spherical Free Air Bursts
The first family of polynomials are established for the range of scaled distances 0.0553 Z
40.0 m/kg1/3 (1.05 r 759) for incident peak overpressure, Ps , incident specific impulse,
I s / W 1/3 , reflected peak overpressure, Pr , reflected specific impulse, I r / W 1/3 , and specific

arrival time, t a / W 1/3 . The ratios of the polynomial to AUTODYN calculations range between
6

For near-field detonations and reflecting surfaces with the geometry of Figure 4-1, the reflected
overpressures and impulses will decrease rapidly from the point of normal reflection due to increasing
scaled distance and angle of incidence. Cormie et al. (2013) extend the utility of the design charts
provided below to address pressures and impulses on reflecting surfaces in the near field.

228

0.90 and 1.10. The polynomial constants are presented in Table 4-1. A range on ratio of (0.95 to
1.05) can be achieved but at the expense of an order higher than 25.
The second set of polynomials are established for subsets of the total range on scaled distance,
namely, 0.0553 Z 0.5 m/kg1/3 and 0.5 Z 40.0 m/kg1/3. For these polynomials, the ratio of
the polynomial to AUTODYN calculations range between 0.95 and 1.05. The polynomial
constants are presented in Table 4-2. The two equations for each parameter share the same value
to four significant figures, and very similar slopes at Z = 0.5 m/kg1/3, as presented in Table 4-3.
Figures 4-2 and 4-3 permit an evaluation of the accuracy of the proposed polynomials as
calculated using the constants of Table 4-2. Data from Needham (2010) are included in Figure 42 for reference. These new polynomial functions are plotted in SI and US units in Figures 4-4
and 4-5, respectively.

229

Table 4-1 Constants of polynomials for incident and normally reflected peak overpressure,
impulse and arrival time for 0.0553 Z 40 m/kg1/3
Parameter
K0
K1
C0
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C11
C12
C13
C14
C15
C16
C17
C18
C19
C20
C21
C22
C23
C24
C25

Ps

I s W 1/3

Pr

I r W 1/3

ta W 1/3

(kPa)

(kPa-ms/kg1/3)

(kPa)

(kPa-ms/kg1/3)

(ms/kg1/3)

-0.1239
0.8705
2.666
-2.769
-0.06247
2.676
0.06411
-4.994
0.181
5.374
-0.4159
-2.715
-0.05814
0.6429
0.208
-0.1394

0.05671
0.8363
2.21
-0.5499
-2.981
2.746
9.522
-13.01
-8.244
12.51
-0.5068
2.939
1.023
-4.748
1.028
-3.996
1.703
0.5049
0.8065
2.006
-1.44
0.6624
-2.118
0.08177
1.678
-0.6299
-0.1429
0.07183

0.1365
0.9839
4.094
-2.439
-1.212
-1.301
3.953
1.69
-4.843
0.2562
2.207
-0.7651
-0.1251
0.0638

0.2044
0.6997
3.105
-2.032
1.36
-1.081
-2.305
3.143
1.151
-2.901
1.003

-0.1674
0.4994
3.351
3.627
-1.24
-3.226
3.144
6.048
-3.093
-5.209

230

Table 4-2 Constants of polynomials for incident and normally reflected peak overpressure, impulse and arrival time for ranges
of scaled distance 0.0553 Z < 0.5 m/kg1/3 and 0.5 Z 40 m/kg1/3

K0
K1
C0
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C11
C12
C13
C14
C15

Ps

I s W 1/3

Pr

(kPa)
Z < 0.5
Z 0.5
-1.888
0.165
-2.603
1.339
4.225
3.266397
0.4076
-1.505
-0.1996
-0.7112
0.2126
-0.02506
0.826
1.842
-0.1719
-1.865
-1.779
0.742
-0.587
-0.114
1.192
0.002556
1.346
0.2694
-1.024
-0.503
0.3686
0.1284
-0.05418

(kPa-ms/kg1/3)
Z < 0.5
Z 0.5
1.206
-0.5596
1.62
1.175
2.30772
1.76
-1.136
-0.6897
1.322
-0.3701
0.7022
-0.1443
-0.2583
1.512
-0.5376
-0.7939
-1.223
-1.814
0.2194
1.639
1.198
-0.2572
-0.8426
0.4388
0.6599
0.1685
-1.029
0.5988
-0.08299

(kPa)
Z < 0.5
Z 0.5
1.613
-0.4784
1.98
0.9222
5.74903
2.251
-1.472
-2.444
0.1963
1.752
1.416
-1.329
-1.999
-1.514
-2.735
4.729
3.556
-3.051
1.131
-2.02
-0.8315
3.52
0.5336
-2.843
-1.676
3.272
-0.2248
-1.233
1.074
-0.6161
-0.2745
1.211
-3.61
2.491

231

I r W 1/3

(kPa-ms/kg1/3)
Z < 0.5
Z 0.5
0.2159
0.8609
0.7506
1.265
3.0934
3.839
-1.892
-2.128
2.003
1.794
2.239
-1.256
0.6488
0.4165
-0.05467
0.0009718

ta W 1/3

(ms/kg1/3)
Z < 0.5
Z 0.5
0.8806
0.267
0.4936
0.7019
0.36193
2.061
1.932
2.148
0.3132
1.98
0.6178
-2.805
1.121
-0.09299

Table 4-3 Values and slopes at Z = 0.5 m/kg1/3 on the polynomials cures plotted using the
two ranges of scaled distance
Parameter

Value
Z < 0.5 m/kg1/3

Slope

Z 0.5 m/kg1/3

Z < 0.5 m/kg1/3

Z 0.5 m/kg1/3

Ps

3.910 MPa

-13.43 MPa-kg1/3/m

-13.65 MPa-kg1/3/m

I s W 1/3

0.1522 MPa-ms/kg1/3

0.1496 MPa-ms/m

0.1803 MPa-ms/m

Pr

26.66 MPa

-108.1 MPa-kg1/3/m

-115.9 MPa-kg1/3/m

I r W 1/3

1.296 MPa-ms/kg1/3

-3.758 MPa-ms/m

-3.620 MPa-ms/m

ta W 1/3

0.1536 ms/kg1/3

0.5130 ms/m

0.5052 ms/m

232

Figure 4-2 Incident peak overpressure impulse and arrival time, spherical free-air bursts

233

Figure 4-3 Reflected peak overpressure and impulse, spherical free-air bursts

234

Figure 4-4 Proposed polynomial charts in SI units, spherical free-air bursts

235

Figure 4-5 Proposed polynomial charts in US units, spherical free-air bursts

236

4.3.2.3 Proposed Polynomials for Hemispherical Free Air Bursts


It is standard practice to increase the charge weight by a factor of 1.8 to transform a spherical
free-air burst to a hemispherical surface burst. The factor is less than idealized value of 2.0 for a
perfectly-reflecting plane, because part of the energy associated with a surface burst is associated
with ground shock, and in forming the crater. To be consistent with current practice, and in the
absence of information to the contrary, the charge weight should be increased by a factor of 1.8
if the charts provided herein are to be used to predict the effects of a hemispherical surface burst.
4.3.3 Sample Results
Table 4-4 presents results of sample calculations using the polynomials with constants per Table 4-2.
Spherical 25 and 1000 kg charges (55 and 2200 lbs, respectively) are considered at scaled distances
of 0.25 and 1.00 m/kg1/3 (0.63 and 2.52 ft/lb1/3, respectively). All of the parameters in the table were
defined previously. Results of the calculations are provided in Figure 4-6.

4.3.4 Use of Design Charts and Considerations of Non-Ideal Explosives


The polynomials and charts proposed in the prior subsection are intended to provide reasonable
predictions for peak overpressure and impulse across a very wide range of scaled distance.
The numerical studies described in Chapters 2 and 3, and in the preceding sections of this
chapter assumed a spherical charge of TNT with a packing density of 1630 kg/m1/3. Blast wave
parameters for other ideal explosives such as RDX, PETN and Composition B and non-ideal
explosives such as ammonium-nitrate-fuel-oil (ANFO) could be estimated by factoring the
charge weight, with different factors often used for overpressure and impulse.
Two other assumptions were made for the numerical simulations, namely, 1) the explosive was a
sphere, and 2) the explosive was detonated centrally. Physical experiments (e.g., Wu et al. 2009)
and numerical simulations (e.g., Sherkar et al. 2010) have shown that charge shape, charge
orientation and point of detonation within the charge can substantially affect overpressure and
impulse, especially in the near field. Importantly, even the detonation of charges of an idealized

237

shape (e.g., sphere, vertical cylinder) will not necessarily produce a uniform wave field along all
radial directions (e.g., Ngo et al. 2014).
The blast analyst must carefully consider the effect of these assumptions when calculating blast
wave pressure histories. It is not practical to address these assumptions by adding variables to the
polynomials presented previously.

238

Table 4-4 Sample calculations for spherical 25 kg and 1000 kg charges and scaled distances of 0.25 m/kg1/3 and 1.00 m/kg1/3
W = 25 kg
1/3

Incident overpressure

Incident impulse

Reflected
overpressure

Reflected impulse

Arrival time

U
Y
Ps (MPa)
U
Y
I s W 1/3
Is (MPa-ms)
U
Y
Pr (MPa)
U
Y
I r W 1/3
Ir (MPa-ms)
U
Y
ta W 1/3
ta (ms)

1/3

Z = 0.25 m/kg

Z = 1.00 m/kg

R = 0.73 m
-0.3208
4.074
11.87
0.2307
2.123
0.1329
0.3885
0.4209
5.192
155.8
-0.2360
3.624
4.208
12.304
0.5834
1.718
0.05229
0.1529

R = 2.92 m
0.1650
3.000
0.9994
-0.5596
2.161
0.1450
0.4239
-0.4784
3.759
5.746
0.8609
2.739
0.5477
1.602
0.2670
2.728
0.5344
1.563

239

W = 1000 kg
Z = 0.25 m/kg1/3
Z = 1.00 m/kg1/3
R = 2.5 m
-0.3208
4.074
11.87
0.2307
2.123
0.1329
1.328
0.4209
5.192
155.8
-0.2360
3.624
4.208
42.08
0.5834
1.718
0.05229
0.5229

R = 10 m
0.1650
3.000
0.9994
-0.5596
2.161
0.1450
1.450
-0.4784
3.759
5.746
0.8609
2.739
0.5477
5.477
0.2670
2.728
0.5344
5.344

(a) Sample results in SI units


Figure 4-6 Use of the proposed polynomials and charts
240

(b) Sample results in US units


Figure 4-6 Use of the proposed polynomials and charts (cont.)
241

SECTION 5
REFLECTED OVERPRESSURE AND IMPULSE AS A FUNCTION OF
ANGLE OF INCIDENCE
5.1 Introduction
Air-blast loading in the far field is typically characterized by a near-instantaneous increase in
overpressure from zero to a peak value followed by an exponential decay to zero: the so-called
positive loading phase. The pressure loading will then fall below ambient (i.e., underpressure),
which is called the negative phase. The positive-phase impulse is the integral of the overpressure
history over the duration of the positive phase. For design using single-degree-of-freedom
procedures (e.g., Biggs 1964, DoD 2008, Cormie et al. 2009, Dusenberry 2010, ASCE 2011), the
exponential decay of overpressure is often replaced by a linear decay to zero to enable the use of
design charts based on triangular loading pulses. The duration of the positive phase for the
linear-decay representation of the loading is chosen so as to preserve the reflected peak
overpressure and reflected total impulse. The negative loading phase is often ignored for design.
Technical guidelines, manuals and textbooks, including TM 5-858-3 (DoA 1984), TM 5-1300
(Department of the Army, Navy and Air Force 1990), Smith and Hetherington (1994), UFC-3340-02 (DoD 2008), Cormie et al. (2009) and Dusenberry (2010), present reflection coefficients
for overpressure. This coefficient is the ratio of the reflected peak overpressure to the incident
peak overpressure and it varies as a function of a) angle of incidence and b) the incident peak
overpressure. Reflected scaled impulse is presented in UFC 3-340-02 as a function of the
incident peak overpressure and the angle of incidence. Together, the reflected peak overpressure
and the reflected impulse can describe a loading function for single-degree-of-freedom
calculations.
Figure 5-1, which is adapted from Norris et al. (1959), illustrates stages in the propagation and
reflection of a shock wave from a free-air burst, which include propagation in free air before
reflection (Figure 5-1a), regular reflection (Figure 5-1b) and irregular or Mach reflection (Figure
5-1c). At an angle of incidence in the range of 40 and 50, the incident and reflected waves
243

merge into a single, reinforced shock front, which is widely known as Mach reflection. The
angle D in Figure 5-1c is the angle of incidence, , associated with the initiation of Mach
reflection.

(a) Before reflection

(b) Regular reflection, D

(c) Mach reflection, D


Figure 5-1 Shock wave reflection phenomena from explosion in free air (adapted from
Norris et al. 1959)
244

Figure 5-2 presents reflection coefficients, Cr , for overpressure from TM 5-858-3 (reproduced in
Smith and Hetherington) and UFC 3-340-02 (superseding TM 5-1300). For overpressure, the
coefficient is two or greater for an angle of incidence of 0 (normal loading) and unity for an
angle of incidence of 90 (side-on loading). The coefficients calculated using the two charts of
Figure 5-2 vary significantly for angles of incidence between 40 and 50, namely, the range
associated with the formation of the Mach stem. Figure 5-3 presents reflected scaled impulse
versus angle of incidence for different values of incident peak overpressure in UFC 3-340-2. The
influence of the Mach stem is not observed in these curves.
Kingery and Bulmash (KB, 1984) developed charts and polynomials for incident and normally
reflected overpressures and impulses as a function of scaled distance, Z, as discussed in Chapter
3. The KB polynomials are implemented in government codes such as CONWEP (Hyde 1992)
and form the basis of the charts in UFC 3-340-02 and TM 5-858-3. The reflection coefficients
are used to compute reflected peak overpressure for non-normal reflection, namely, greater
than 0. The reflected impulse for a non-zero angle of incidence is calculated using the UFC 3340-02 charts for reflected scaled impulse and the charge weight.
Chapter 3 demonstrated that the KB polynomials underestimate incident and normally reflected
overpressures and incident impulse in the near field. Chapter 4 proposed new equations and
charts for incident and normally reflected peak overpressure and impulse for 0.0553 Z 40
m/kg1/3 (0.139 Z 100 ft/lb1/3), using numerical data reported in Chapter 3. Figures 5-4 and 5-5
present the charts proposed in Chapter 4 and these are used in the presentation that follows.
This chapter reconciles the differences between overpressure reflection coefficients in UFC 3340-02 and TM 5-858-3 to enable accurate prediction of reflected peak overpressure for all
angles of incidence and incident peak overpressure in the range from 0.016 MPa to 20 MPa. The
CFD tool, AUTODYN (ANSYS 2013a), which is verified and validated in Chapter 2, is used for
the analysis. Reflected scaled impulse is also studied to a) evaluate the charts in UFC 3-340-02
that are reproduced in Figure 5-3, and b) provide guidance on computational fluid dynamics
(CFD) modeling of reflected impulse as a function of angle of incidence.

245

(a) TM 5-858-3 (DoA 1984)

(b) UFC 3-340-02 (DoD 2008)


Figure 5-2 Reflection coefficient versus angle of incidence
246

(a) 10 Ps 7000 psi (0.069 Ps 48.3 MPa)

(b) 0.7 Ps 1500 psi (0.0048 Ps 10.3 MPa)


Figure 5-3 Reflected scaled impulse versus angle of incidence (UFC 3-340-02 (DoD 2008))
247

Figure 5-4 Incident, Ps , and normally reflected, Pr , peak overpressures developed by


Kingery and Bulmash (1984) and in Chapter 4

248

Figure 5-5 Incident, I s /W1/3, and normally reflected, I r /W1/3, scaled impulses developed by
Kingery and Bulmash (1984) and in Chapter 4

249

5.2 Regular and Mach Reflections


Figures 5-2a and 5-2b show the value of the reflection coefficient, Cr , decreasing with
increasing angle of incidence, , until the angle reaches approximately 40o, which is the region
of regular reflection as illustrated in Figure 5-1b. At greater angles, the reflected wave, which is
passing through heated and compressed air, and hence propagating faster than the incident wave,
catches and merges with the incident wave at D , as shown in Figure 5-1c. The merging of
the wave fronts gives rise to the Mach stem illustrated in the figure. For angles of incidence
greater than that associated with the formation of a Mach stem, the reflection is termed a Mach,
or irregular, reflection. The significant differences between Figures 5-2a and 5-2b are associated
with Mach reflection. Figure 5-6, which is adapted from Smith and Hetherington (1994),
provides a simple illustration of regular and Mach reflection. The angle of reflection, , equals
the angle of incidence, , for weak (acoustic) shocks. For strong shocks, Griffith and Bleakney
(1954) reported that the ratio : is greater than 1 and the ratio increases with the intensity of
the shock.

(a) Regular reflection

(b) Mach reflection

Figure 5-6 Regular and Mach reflections (e.g., Smith and Hetherington 1994)
The regions of regular and Mach reflections can be established as a function of the Mach number
of the incident shock front (shock front velocity divided by sound speed in air) and the angle of
incidence. Figure 5-7 (adapted from Ivanov et al. 2001) illustrates the relationships between
these reflections, the Mach number, M, and the angle of incidence, , for an ideal gas. The
angles, N and D , correspond to the von Neumann and detachment criteria, respectively. For
< N , only regular reflections are possible, and for > D , only Mach reflections are possible.
For M less than 2.2, N and D are identical. For M > 2.2, the difference between N and D
250

increases with Mach number. Both regular and Mach reflections are possible in the domain

N D . In this domain, which is described as the dual solution domain (DSD), reflections
can transition between regular and Mach as a result of unsteadiness in the flow (Mouton 2007).
For a steady flow, regular reflection is likely to occur in the DSD and this is assumed herein,
namely, D , is associated with Mach reflection.

Figure 5-7 Dual solution domain (Ivanov et al. 2001)


For incident peak overpressures less than 0.34 MPa (50 psi), the Mach reflection angle increases
as the incident overpressure decreases, as shown in Figure 5-8 and as reported by Courant and
Friedrichs (1948) and Kinney and Graham (1985). This observation is consistent with the charts
in TM 5-858-3 and UFC 3-340-02, which are reproduced in Figure 5-2.
TM 5-858-3 assumes the angle of incidence for Mach reflection to be approximately 40o for
incident peak overpressures greater than approximately 0.34 MPa (50 psi). The assumption is
appropriate for an ideal gas with a specific heat ratio of 1.4 (e.g., Courant and Friedrichs 1948,
Kinney, 1962, Kinney and Graham 1985, Schwer 2008). Figure 5-8 (adapted from Kinney and
Graham 1985) presents the relationship between the angle of incidence associated with the
formation of a Mach stem (i.e., onset of Mach reflection) and the Mach number of the shock
front. For Mach number greater than 2.0, the angle of incidence is approximately 40; which is
seen in Figure 5-2a for incident overpressures greater than 50 psi (0.34 MPa). An incident peak

251

overpressure of 0.34 MPa corresponds to a shock front velocity of approximately Mach 2 (see
Figure 5-9).

Figure 5-8 Angle of incidence for Mach reflection as a function of Mach number for an
ideal gas (Kinney and Graham 1985)

Figure 5-9 Incident overpressure as a function of Mach number for an ideal gas (Kinney
and Graham 1985)
In UFC 3-340-02, the effect of Mach reflection is observed in the range 45 < < 50 for
incident peak overpressures of 2.8 MPa (400 psi) and greater, which corresponds approximately
to Cr > 7 at = 0 (see Figure 5-2b). The angle of incidence associated with Mach reflection
increases with a decrease in the specific heat ratio that accompanies higher incident
overpressures (Schwer 2008). Schwer computed angles of incidence for Mach reflection, as a
function of Mach number from 3 to 18, and reported that for an ideal gas with a specific heat
252

ratio a) of 1.4, the angle was constant at 39.9o between Mach 3 and Mach 18 (supporting the
value of approximately 40 in TM 5-858-3), and b) of 1.2, the angle increased from 42 at Mach
3 to 46 at Mach 18 (supporting the increase in angle with increasing incident overpressure seen
in the UFC).
The effect of Mach reflection is not observed in the UFC 3-340-02 charts of reflected scaled
impulse that are reproduced in Figure 5-3, for the reason given below.

5.3 A Numerical Model for Calculating Reflection Coefficients


5.3.1 Modeling and Domains
Reflection coefficients are calculated by simulating the detonation of a 960 kg spherical charge
of TNT in free air, as described in Chapters 3 and 4. The expansion of the detonation products is
modeled using the JWL Equation of State (EOS) in 1D multi-material Euler-Godunov grids (see
Section 2.2.1). The 1D simulation proceeds until the shock wave reaches a reflecting boundary.
The 1D results are then mapped into 2D Euler-FCT grids, as illustrated in Figure 3-7. After the
detonation product has expanded to 10 times its original volume, standard practice replaces the
JWL EOS with the ideal gas EOS to avoid possible numerical instability. The Dobratz and
Crawford (1985) values of the parameters for the JWL EOS, presented in Table 2-1, are used for
the simulations described below.
Calculations are performed for one charge weight only. The results of simulations of incident
and reflected peak overpressure in Chapter 3 for three very different charge masses (18735, 960
and 23 kg) showed no influence of weight for a given scaled distance and so calculations are
performed using only one charge mass.
In the model depicted in Figure 5-10, axial symmetry is used at the lower boundary and mirror
symmetry is used at the left boundary; 1/4 of the explosive and air domains are analyzed. The
mirrored-symmetric boundary is a perfectly reflecting surface. The monitoring location is varied
by angle, , and distance, R, to calculate reflected peak overpressures as a function of the angle
of incidence. The length and height of the air domain are R cos and ( R cos + R/4),

253

respectively. The length of R/4 is chosen to eliminate partial reflection7 of overpressure from the
horizontal transmitting (upstream) boundary on the reflected peak overpressure at the monitoring
location. Numerical simulations are performed for fourteen scaled distances, 0.16, 0.2, 0.26, 0.3,
0.4, 0.6, 0.8, 1.2, 1.6. 2.4, 3.2, 4, 6 and 8 m/kg1/3 (0.4, 0.5, 0.65, 0.75, 1, 1.5, 2, 3, 4, 6, 8, 10, 15
and 20 ft/lb1/3, respectively) and fifteen angles of incidence, 0, 10, 20, 30, 40, 42.5, 45,
47.5, 50, 52.5, 55, 57.5, 60, 70 and 80, to establish the relationships between overpressure
reflection coefficient, angle of incidence and incident peak overpressure. Smaller intervals are
used in the range 40 60 to capture the effect of Mach reflection. For an angle of 90, the
reflected overpressure is identical to the incident overpressure. The scaled distances of 0.16, 0.2,
0.26 and 0.3 m/kg1/3 are not simulated for = 80 because the perpendicular distance from the
reflecting surface to the center of the charge is less than the charge radius; see Figure 5-10. For Z
= 0.3 m/kg1/3 (or R = 2.96 m) and = 80, the length of the air domain, R cos = 0.514 m is
smaller than the charge radius (= 0.520 m for TNT weight of 960 kg and packing density of 1630
m/kg1/3). The smallest scaled distance of 0.16 m/kg1/3 enables calculation of reflection
coefficients with angles of incidence up to 70. (For R cos = 0.520 m and = 70 in Figure 510, R = 1.52 m, which corresponds to Z = 0.154 m/kg1/3.) The maximum scaled distance of 8
m/kg1/3 used here produces an incident peak overpressure of 0.015 MPa, approximately 1/7 of
ambient pressure at sea level (= 0.1 MPa), which is a practical lower limit for the blast-resistant
design of structural components

A transmitting boundary does not perfectly pass outgoing pressure. A small fraction of the outgoing
wave is reflected due to numerical approximations. Although the amplitudes of the partially reflected
waves are generally small, they can increase both the reflected peak overpressure and impulse if the
distance of the transmitting boundary from the monitoring location is very small. A distance of R/4 is
sufficient to eliminate any significant effect of wave reflection; see Section 2.5 for details.

254

Figure 5-10 2D numerical model of air for calculating reflection coefficients


5.3.2 Cell Size
One advantage of the remapping procedure described previously is that fine meshes can be used
for 1D simulations and coarser meshes used for 2D simulations. A cell size of R/1600 is used for
1D radial analysis of incident peak overpressure, where R is the distance to the monitoring
location from the center of the charge. Results are then mapped into a 2D domain, with a cell
size of R/800. These cell sizes for 1D and 2D analysis are smaller than 0.002 times the distance
to the monitoring location (R/500) deemed appropriate in Chapter 3 for simulations of incident
and reflected peak overpressures at Z 8 m/kg1/3.

5.4 Numerical Models for Calculating Reflected Impulses


5.4.1 Modeling and Domains
Reflected impulse is calculated for the 960 kg spherical charge of TNT and scaled distances of
0.16 to 8 m/kg1/3, used in the analysis of reflection coefficient in Section 5.3, but for fewer (nine)
angles of incidence of 0, 10, 20, 30, 40, 50, 60, 70 and 80 because preliminary studies
indicated that Mach reflection had no significant effect on reflected scaled impulse for angles of
incidence in the range 40 50. (This is confirmed in Section 5.8 using reflected
255

overpressure and scaled impulse histories for Z = 0.16, 0.40 and 1.2 m/kg1/3 with = 30, 40,
45, 50 and 55.) For an angle of 90, the reflected impulse for a given scaled distance is
identical to the incident impulse. The modeling of the expansion of the detonation products (e.g.,
JWL EOS for TNT, 1D to 2D mapping) is similar to that described in the previous section.
Three AUTODYN 2D models of air are constructed to calculate reflected impulse as a function
of angle of incidence; see Figure 5-11. A quarter of the charge is modeled with axial symmetry at
the lower boundary. Model 1 is similar to that shown in Figure 5-10. The upper transmitting
boundary in Figure 5-11a should prevent reflection of pressure waves that would affect the
calculation of reflected impulse. However, as noted in Chapter 2, partial reflection is possible
from a transmitting boundary, and the boundary must be located a sufficient distance from the
monitoring locations. (In Chapter 2, the boundary was set 149 mm above the uppermost
4 in Figure 2-49a.) For these calculations, the height of the 2D domain is
monitoring location:

calculated by adding the shortest distance from the detonation point to the right reflecting
boundary (= R cos ) to the height of the monitoring location, which is sufficient to make
accurate calculations of reflected impulse. The left transmitting boundary in Figure 5-11a should
also prevent the reflection of pressure waves between the vertical boundaries. Energy will not be
lost across this boundary in the initial phase of the expansion because the shock front propagates
radially away from the source and thus away from this boundary. However, the wave that is
reflected from the right boundary may be partially reflected from the left transmitting boundary,
which would then increase the impulse if the partially reflected wave arrives at the monitoring
location on the right boundary during the positive phase. This is more likely to occur for higher
angles of incidence with a relatively small horizontal 2D domain. Models 2 and 3 increase the
horizontal length of the 2D domain of Model 1 to evaluate the effect of location of the left
transmitting boundary on reflected impulse for large angles of incidence. Models 1, 2 and 3 have
horizontal dimensions of R cos , 1.5 R cos and 2 R cos , respectively, as shown in Figure 511. Figure 5-12 illustrates Models 1, 2 and 3 for an angle of incidence of 80, where reflected
scaled impulse is calculated at point 1. The use of the three models is discussed further in Section
5.4.3, noting that analysis of Model 3 is computationally more expensive than analysis of Model
1.

256

(a) Model 1

(b) Model 2

(c) Model 3
Figure 5-11 2D numerical models of air for calculating reflected impulses

257

(a) Model 1

(b) Model 2

(c) Model 3

Figure 5-12 Models 1, 2 and 3 in AUTODYN for angle of incidence of 80


5.4.2 Cell Size
One-dimensional radial analysis is performed using a cell size of R/1600 as discussed in Section
5.3. The 1D results are then mapped into a 2D domain. Two cell sizes, R/800 and R/400, are
considered for the 2D analysis because the calculation of reflected impulse converges for larger
cell sizes than reflected peak overpressure, as reported in Section 3.7. Figure 5-13 presents
normally reflected overpressure and impulse histories calculated using Model 1 for cell sizes of
R/400 and R/800 and for Z = 0.16, 0.40 and 1.2 m/kg1/3. The results are essentially independent
of cell size at each scaled distance. Since the use of the smaller cell size (R/800) is
computationally more expensive, the larger cell size (R/400) is used to model the 2D domain for
calculating reflected scaled impulse.
258

Reflected overpressure history

Reflected impulse history

(a) Z = 0.16 m/kg1/3

(b) Z = 0.40 m/kg1/3

(c) Z = 1.2 m/kg1/3


Figure 5-13 Mesh sensitivity study for simulating reflected impulse; Model 1
259

5.4.3 Effect of Transmitting Boundary


The effect of the location of the left transmitting boundary on the reflected impulse calculation is
examined using Models 1, 2 and 3 for Z = 0.16, 0.40 and 1.2 m/kg1/3. Figure 5-14 presents results
calculated using the cell size of R/400 for Z = 0.16 m/kg1/3 with = 30, 40, 50, 60 and 70.
The reflected impulses for Models 1, 2 and 3 are very similar for = 30 and 40. For = 50,
60 and 70, the reflected impulses for Model 1 are significantly greater than those for Models 2
and 3 due to partial reflection of the shock waves from the left transmitting boundary. This
outcome is similar for Z = 0.40 and 1.2 m/kg1/3 with 70, as shown in Figures 5-15 and 5-16,
respectively. Results for Models 2 and 3 are essentially identical.
The influence of the location of the left transmitting boundary is greater for Z = 0.16 m/kg1/3 than
for Z = 0.40 and 1.2 m/kg1/3 because a) the shock waves that are partially reflected from the left
(transmitting) boundary propagate faster at smaller scaled distances, and b) the effect of the
expanding detonation products can increase the positive phase duration for small values of scaled
distance as demonstrated in Chapter 2 and discussed further in Section 5.6. Accordingly, for Z =
0.16, 0.20, 0.26 and 0.30 m/kg1/3, Model 1 is used for 0 40 and Model 2 is used for >
40. For Z 0.40 m/kg1/3, Model 1 is used for 0 60 and Model 2 is used for > 60.
Model 3 is not used for subsequent calculations.

260

Reflected overpressure history

Reflected impulse history

(a) = 30

(b) = 40

(c) = 50
Figure 5-14 Effect of transmitting boundary on calculation of reflected impulse for Z = 0.16
m/kg1/3
261

Reflected overpressure history

Reflected impulse history

(d) = 60

(e) = 70
Figure 5-14 Effect of transmitting boundary on calculation of reflected impulse for Z = 0.16
m/kg1/3 (cont.)

262

Reflected overpressure history

Reflected impulse history

(a) = 40

(b) = 50

(c) = 60
Figure 5-15 Effect of transmitting boundary on calculation of reflected impulse for Z = 0.40
m/kg1/3
263

Reflected overpressure history

Reflected impulse history

(d) = 70

(e) = 80
Figure 5-15 Effect of transmitting boundary on calculation of reflected impulse for Z = 0.40
m/kg1/3 (cont.)

264

Reflected overpressure history

Reflected impulse history

(a) = 40

(b) = 50

(c) = 60
Figure 5-16 Effect of transmitting boundary on calculation of reflected impulse for Z = 1.2
m/kg1/3
265

Reflected overpressure history

Reflected impulse history

(d) = 70

(e) = 80
Figure 5-16

Effect of transmitting boundary on calculation of reflected impulse for Z =

1.2 m/kg1/3 (cont.)

266

5.5 Incident and Normally Reflected Overpressures


Incident, Ps , and reflected, Pr , peak overpressures for different values of scaled distance and
normal incidence, = 0, are calculated using cell sizes of R/1600 (1D) and R/800 (2D). Results
are presented in Figure 5-17, together with data from polynomials per Table 4-2 of Chapter 4 and
from UFC 3-340-02. Values are reported in Table 5-1, noting that the polynomials were derived
from AUTODYN analysis using finer meshes than those used in this chapter. Table 5-2 presents
ratios of the AUTODYN calculations to a) the predictions of the polynomials of Chapter 4, and b)
the UFC 3-340-02. (The ratio of AUTODYN-to-polynomials ranges between 0.9 and 1.1
because 10% was the limit set for a converged cell size, as noted in Sections 3.6 and 3.7.) For
the scaled distances considered here (Z 0.16 m/kg1/3), the AUTODYN-calculated incident peak
overpressures are very close to the values reported in UFC 3-340-02. The calculated normally
reflected peak overpressures are also similar to those of UFC 3-340-02 for Z 0.4 m/kg1/3 (1
ft/lb1/3). For Z < 0.4 m/kg1/3, the AUTODYN-calculated normally reflected peak overpressure is
much greater than that in UFC 3-340-02; see Figure 5-4, which is reproduced from Figures 4-5
and 4-6. Reflected peak overpressures for angles of incidence greater than 0o (normal reflection)
are studied in the following section.

5.6 Reflected Overpressure as a Function of Angle of Incidence


Reflection coefficients, Cr , for different angles of incidence, , calculated using AUTODYN
are presented in Figures 5-18a, 5-18b8, 5-18c8 and Table 5-3. The effect of Mach reflection is
identified by local increases in the plots for 40 < < 50 in all these curves: similar to UFC 3340-02 but dissimilar from TM 5-858-3 for incident peak overpressures greater than 2.8 MPa
(400 psi).

Figures 5-18b and 5-18c present the same data at Figure 5-18a but in a format that is simpler to interpret.

267

Figure 5-17 Incident and normally reflected peak overpressures

268

Table 5-1 Incident and reflected overpressures for normal incidence


AUTODYN (ANSYS 2013a)
Z

Proposed polynomials (Chapter 4)

Pr

Ps

UFC 3-340-02 (DoD 2008)

Pr

Ps

Pr

Ps

m/kg1/3

ft/lb1/3

MPa

Psi

MPa

psi

MPa

psi

MPa

psi

MPa

psi

MPa

psi

0.16

0.40

19.7

2860

466

67570

19.7

2863

495

71810

19.6

2840

214

31000

0.20

0.50

15.1

2190

263

38100

15.7

2278

271

39260

15.3

2220

158

22900

0.26

0.65

10.9

1580

143

20700

11.3

1633

141

20450

11.0

1590

106

15400

0.30

0.75

8.86

1280

94.7

13700

9.26

1344

96.0

13920

9.00

1300

83.3

12100

0.40

1.0

5.62

815

45.3

6570

5.84

847

43.2

6269

5.75

833

48.8

7070

0.60

1.5

2.77

402

19.2

2790

2.84

412

18.5

2689

2.74

398

19.8

2870

0.80

2.0

1.62

235

10.3

1490

1.63

237

10.3

1493

1.52

221

9.38

1360

1.2

3.0

0.640

92.8

3.19

463

0.651

94.4

3.31

480

0.620

89.9

2.93

425

1.6

4.0

0.310

45.0

1.25

182

0.324

47.0

1.31

191

0.322

46.7

1.24

180

2.4

6.0

0.125

18.1

0.366

53.1

0.127

18.4

0.393

57.0

0.130

18.9

0.387

56.1

3.2

8.0

0.073

10.5

0.184

26.7

0.071

10.2

0.191

27.8

0.072

10.4

0.184

26.7

4.0

10

0.048

7.02

0.116

16.8

0.047

6.79

0.118

17.2

0.047

6.75

0.110

16.0

6.0

15

0.025

3.63

0.055

7.94

0.024

3.50

0.056

8.09

0.023

3.37

0.051

7.32

8.0

20

0.016

2.39

0.031

4.54

0.016

2.27

0.035

5.05

0.015

2.19

0.032

4.65

269

Table 5-2 Ratios of AUTODYN calculations to proposed polynomial values and UFC 3-34002 (DoD 2008) predictions
AUTODYN/polynomials

AUTODYN/UFC 3-340-02

m/kg1/3

ft/lb1/3

Ps

Pr

Ps

Pr

0.16

0.40

0.94

0.94

1.01

2.34

0.20

0.50

0.96

0.97

0.99

1.66

0.26

0.65

0.97

1.01

0.99

1.34

0.30

0.75

0.95

0.98

0.98

1.13

0.40

1.0

0.96

1.05

0.98

0.93

0.60

1.5

0.98

1.04

1.01

0.97

0.80

2.0

0.99

1.00

1.06

1.10

1.2

3.0

0.98

0.96

1.03

1.09

1.6

4.0

0.96

0.95

0.96

1.01

2.4

6.0

0.98

0.93

0.96

0.95

3.2

8.0

1.03

0.96

0.98

0.99

4.0

10

1.03

0.98

1.01

1.01

6.0

15

1.04

0.98

1.05

1.04

8.0

20

1.05

0.90

1.05

1.01

270

Table 5-3 Reflection coefficients as a function of angle of incidence, , and scaled distance, Z, calculated using AUTODYN
Angle of incidence

Z
m/kg1/3

ft/lb1/3

10

20

30

40

42.5

45

47.5

50

52.5

55

57.5

60

70

0.16

0.40

23.7

23.0

21.3

15.4

11.5

11.4

9.0

7.9

14.5

8.3

7.7

8.3

7.8

2.2

1.0

0.20

0.50

17.4

17.4

17.2

13.9

9.3

7.7

7.9

10.9

7.5

6.5

5.1

4.6

5.0

2.7

1.0

0.26

0.65

13.2

12.6

11.3

9.9

6.9

6.8

5.7

5.9

7.3

5.8

4.4

4.2

4.3

2.8

1.0

0.30

0.75

10.7

10.4

10.1

8.4

6.7

6.7

5.7

5.3

6.7

5.6

4.5

3.6

3.5

2.8

1.0

0.40

1.0

8.1

7.8

7.2

6.6

6.5

6.3

5.5

4.9

6.3

4.8

4.3

3.7

3.6

2.5

2.1

1.0

0.60

1.5

6.9

6.7

6.5

6.0

5.9

5.7

4.7

4.4

4.9

5.1

3.3

3.3

3.2

2.4

1.8

1.0

0.80

2.0

6.3

6.2

6.1

5.7

5.6

5.7

4.4

4.0

3.8

3.7

3.6

3.0

2.8

2.3

1.6

1.0

1.2

3.0

5.0

4.9

4.8

4.5

4.7

4.9

3.7

3.1

2.8

2.6

2.5

2.5

2.5

2.4

1.8

1.0

1.6

4.0

4.0

4.0

3.9

3.8

4.0

4.2

3.8

3.0

2.7

2.6

2.5

2.5

2.4

2.3

1.8

1.0

2.4

6.0

2.9

2.9

2.9

2.9

3.1

3.2

3.4

3.3

2.9

2.7

2.5

2.4

2.3

2.0

1.6

1.0

3.2

8.0

2.5

2.5

2.6

2.5

2.6

2.7

2.8

3.0

3.1

3.0

2.7

2.6

2.4

2.0

1.7

1.0

4.0

10

2.4

2.4

2.4

2.4

2.4

2.5

2.5

2.6

2.8

3.0

3.0

2.8

2.6

2.0

1.7

1.0

6.0

15

2.2

2.2

2.2

2.2

2.2

2.2

2.2

2.2

2.3

2.4

2.6

2.8

2.9

2.2

1.7

1.0

8.0

20

1.9

2.1

2.1

2.1

2.0

2.0

2.1

2.1

2.1

2.2

2.2

2.4

2.5

2.4

1.7

1.0

271

80

90

For incident peak overpressures of 0.016, 0.025, 0.048, 0.073 and 0.13 MPa (Z = 8.0, 6.0, 4.0,
3.2 and 2.4 m/kg1/3; see Figure 5-18), the angle of incidence associated with Mach reflection is
greater than 40o because the angle increases with decreasing incident peak overpressure below
0.31 MPa (or Mach number less than 2.0) per Figures 5-8 and 5-9.
The angles of incidence for Mach reflection for Ps = 0.31, 0.64 and 1.6 MPa (associated with Z
= 1.6, 1.2 and 0.8, respectively, and the curves with Cr = 4.0, 5.0 and 6.3 for normal reflection,
respectively, in Figure 5-18) are close to 40: a result that is in good agreement with Figure 5-8.

(a) Scaled distances, Z, between 0.16 and 8.0 m/kg1/3 (1 m/kg1/3 =2.52 ft/lb1/3; 1 MPa = 145 psi)
Figure 5-18 Overpressure reflection coefficients as a function of angle of incidence

272

(b) Scaled distances, Z, between 0.16 and 0.6 m/kg1/3 (1 m/kg1/3 =2.52 ft/lb1/3; 1 MPa = 145 psi)
Figure 5-18 Overpressure reflection coefficients as a function of angle of incidence (cont.)

273

(c) Scaled distances, Z, between 0.6 and 8.0 m/kg1/3 (1 m/kg1/3 =2.52 ft/lb1/3; 1 MPa = 145 psi)
Figure 5-18 Overpressure reflection coefficients as a function of angle of incidence (cont.)
The angles of incidence associated with Mach reflection for Ps 2.8 MPa (associated with Z
0.6 m/kg1/3 and the curves with Cr 7 for normal reflection) are between 45 and 50, and
similar to the values in UFC 3-340-02 (see Figure 5-2b). For these higher incident overpressures,
the specific heat ratio decreases and the Mach reflection angle increases, which is consistent with
the observations of Schwer.
To understand the influence of the expanding detonation products on reflection coefficients for Z
0.8 m/kg1/3, Figure 5-19 presents histories of reflected overpressure, pr (t ) , incident
overpressure, p s (t ) , density, s (t ) , and flow (particle) velocity, s (t ) , for Z = 0.16, 0.4, 0.8 and
274

1.2 m/kg1/3 for a) normal reflection, and b) the angle of incidence for which the peak
overpressure is the greatest in the region of Mach reflection. The histories are amplitude scaled
to facilitate presentation in one figure. For Z = 0.16 m/kg1/3, the arrival times of the shock front
and detonation products are approximately t = 0.26 and 0.29 msec, respectively, as shown in
Figure 5-19a. The differences between the arrival times of the incident and reflected
overpressures are caused by the use of different numerical procedures: 1D and Euler-Godunov
grid for the incident overpressure, and 2D and Euler-FCT grid for the reflected overpressure. The
reflected peak overpressures in Figure 5-19a are associated with the expanding detonation
products and not the shock fronts. Two peaks are predicted in the normal reflection overpressure
history and identified by (*) in the Figure 5-19a. The first is associated with the arrival of the
front of the detonation products. The second is associated with the peak particle velocity,
identified by (#) in the figure, at the tail of the expanding detonation products (the second *
should align in time with # but is offset for the reason noted above). The reflected overpressure is
a function of the static overpressure (twice the incident overpressure) and the dynamic
overpressure (related to the air density and particle velocity): the greater density and higher
velocity of the detonation products produces a larger reflected overpressure for this value of
scaled distance than at the shock front. The reflected overpressures histories in Figure 5-19a
fluctuate due to the influence of the expanding detonation products.
For Z = 0.4 m/kg1/3, the reflected peak overpressure is associated with the arrival of the shock
front at t = 1.0 msec, as seen in Figure 5-19c. The detonation products arrive at monitoring
location at approximately t = 1.25 msec. The arrival times of the incident and reflected shock
fronts are again different for the reason given above. The expanding detonation products lead to
fluctuations in the overpressure histories, similar to that observed for Z = 0.16 m/kg1/3.
For Z = 0.8 m/kg1/3, the influence of the detonation products on the overpressure histories is
negligible, as seen in Figure 5-19e: the shapes of the overpressure histories are similar to the
Friedlander curve, namely, a smooth exponential decay from the peak overpressure to ambient
(see Equation 1.1). The Friedlander overpressure history is achieved for incident and reflected
peak overpressures of Z > 0.83 m/kg1/3, as seen in Figure 5-19g for Z = 1.2 m/kg1/3.

275

(a) Z = 0.16 m/kg1/3, = 0

(b) Z = 0.16 m/kg1/3, = 50

(c) Z = 0.40 m/kg1/3, = 0

(d) Z = 0.40 m/kg1/3, = 50

(e) Z = 0.8 m/kg1/3, = 0

(f) Z = 0.8 m/kg1/3, = 42.5

Figure 5-19 Reflected overpressure histories; Z = 0.16, 0.40, 0.80 and 1.2 m/kg1/3 (1 m/kg1/3
=2.52 ft/lb1/3; 1 MPa = 145 psi)
276

(g) Z = 1.2 m/kg1/3, = 0

(h) Z = 1.2 m/kg1/3, = 42.5

Figure 5-19 Reflected overpressure histories; Z = 0.16, 0.40, 0.80 and 1.2 m/kg1/3 (1 m/kg1/3
=2.52 ft/lb1/3; 1 MPa = 145 psi) (cont.)
Figure 5-20 enables a comparison of data from the AUTODYN simulations and empirical results
from TM 5-858-3 and UFC-3-340-02. The overpressure reflection coefficients obtained from the
AUTODYN simulations for Z < 0.4 m/kg1/3 (1 ft/lb1/3) are greater than those reported in both
UFC 3-340-02 and TM 5-858-3 for angles of incidence less than 40o, as shown in panels a, b, c
and d of Figure 5-20 (Z = 0.16, 0.20, 0.26 and 0.30 m/kg1/3, respectively) due to the effects of the
expanding detonation products on the reflected overpressure histories. For Z 0.4 m/kg1/3 and
< 40, the overpressure reflection coefficients calculated with AUTODYN are in reasonable
agreement with the values reported in both UFC 3-340-02 and TM 5-858-3, as shown in panels e,
f, g, h, i, j, k, l, m and n of Figure 5-20. For Z 0.4 m/kg1/3 and > 40, the AUTODYN-based
overpressure reflection coefficients are similar to the values reported in UFC 3-340-02 but are
quite different from those of TM 5-858-3.

277

(a) Z = 0.16 m/kg1/3 (0.4 ft/lb1/3)

(b) Z = 0.2 m/kg1/3 (0.5 ft/lb1/3)

(c) Z = 0.26 m/kg1/3 (0.66 ft/lb1/3)

(d) Z = 0.3 m/kg1/3 (0.75 ft/lb1/3)

(e) Z = 0.4 m/kg1/3 (1 ft/lb1/3)

(f) Z = 0.6 m/kg1/3 (1.5 ft/lb1/3)

Figure 5-20 Overpressure reflection coefficients with angle of incidence from TM 5-858-3
(DoA 1984), UFC 3-340-02 (DoD 2008), numerical analyses by AUTODYN; P s
in units of MPa

278

(g) Z = 0.8 m/kg1/3 (2 ft/lb1/3)

(h) Z = 1.2 m/kg1/3 (3 ft/lb1/3)

(i) Z = 1.6 m/kg1/3 (4 ft/lb1/3)

(j) Z = 2.4 m/kg1/3 (6 ft/lb1/3)

(k) Z = 3.2 m/kg1/3 (8 ft/lb1/3)

(l) Z = 4 m/kg1/3 (10 ft/lb1/3)

Figure 5-20 Overpressure reflection coefficients with angle of incidence from TM 5-858-3
(DoA 1984), UFC 3-340-02 (DoD 2008), numerical analyses by AUTODYN;
P s in units of MPa (cont.)
279

(m) Z = 6 m/kg1/3 (15 ft/lb1/3)

(n) Z = 8 m/kg1/3 (20 ft/lb1/3)

Figure 5-20 Overpressure reflection coefficients with angle of incidence from TM 5-858-3
(DoA 1984), UFC 3-340-02 (DoD 2008), numerical analyses by AUTODYN;
P s in units of MPa (cont.)
For reflection coefficients to be useful for design, the incident and reflected pressure histories
should follow a standard shape, such as the Friedlander curve, with the same positive phase
duration. TM 5-858-3 and UFC 3-340-02 present data (see Figure 5-2) for incident peak
overpressures up to 7000 psi (48.3 MPa) and 5000 psi (34.5 MPa), respectively. Herein
simulations were performed for incident peak overpressures up to 20 MPa (2900 psi). As seen in
Figure 5-19, at Z = 0.4 m/kg1/3 and Ps = 6 MPa (870 psi), the incident and reflected overpressure
histories are very different from the Friedlander curve, and values of reflected peak overpressure
and scaled impulse alone are is of no practical value for blast-resistant design: CFD analysis is
needed to adequately characterize incident and reflected pressure histories. A practical lower
limit on the use of reflection coefficients is Z = 0.8 m/kg1/3.

5.7 Incident and Normally Reflected Impulses


Incident, I s /W1/3, and normally reflected, I r /W1/3, scaled impulses are calculated using cell sizes
of R/1600 (1D) and R/400 (2D). Results are presented in Figure 5-21, together with data from the
polynomials per Table 4-2 of Chapter 4 and from UFC 3-340-02. Values are provided in Table
5-4. Table 5-5 presents ratios of the AUTODYN calculations to a) the polynomial predictions of
Chapter 4, and b) the values calculated using UFC 3-340-02. (The ratio of AUTODYN-to-

280

polynomial ranges between 0.98 and 1.07.) The AUTODYN calculations and UFC 3-340-02based predictions of incident scaled impulse differ by 10% to 20%, which is consistent with
observations in Chapters 3 and 4. The AUTODYN-calculated normally reflected scaled impulses
are very similar to those reported in UFC 3-340-02.

Figure 5-21 Incident and normally reflected scaled impulses


281

Table 5-4 Incident and normally reflected scaled impulses in SI units of MPa-ms/kg1/3 and US units of psi-ms/lb1/3
AUTODYN (ANSYS 2013a)

I s W 1/3

Proposed polynomials (Chapter 4)

I r W 1/3

I s W 1/3

UFC 3-340-02 (DoD 2008)

I r W 1/3

I s W 1/3

I r W 1/3

m/kg1/3

ft/lb1/3

SI

US

SI

US

SI

US

SI

US

SI

US

SI

US

0.16

0.40

0.275

30.7

9.83

1096

0.258

28.7

9.91

1104

0.317

35.3

9.12

1016

0.20

0.50

0.182

20.3

6.37

710

0.170

19.0

6.43

717

0.226

25.2

6.18

688

0.26

0.65

0.138

15.4

3.86

430

0.130

14.4

3.91

436

0.168

18.7

3.99

444

0.30

0.75

0.131

14.6

2.96

330

0.124

13.8

3.01

335

0.151

16.8

3.00

335

0.40

1.0

0.140

15.6

1.80

201

0.135

15.0

1.83

204

0.137

15.2

2.03

226

0.60

1.5

0.173

19.3

1.01

112

0.177

19.8

1.02

113

0.155

17.2

1.12

125

0.80

2.0

0.183

20.4

0.709

79.1

0.177

19.7

0.712

79.3

0.196

21.8

0.752

83.8

1.2

3.0

0.118

13.1

0.454

50.6

0.120

13.3

0.444

49.5

0.149

16.6

0.442

49.3

1.6

4.0

0.092

10.3

0.329

36.7

0.092

10.2

0.319

35.5

0.113

12.6

0.309

34.4

2.4

6.0

0.069

7.64

0.199

22.2

0.068

7.60

0.198

22.1

0.078

8.73

0.190

21.2

3.2

8.0

0.054

6.01

0.138

15.4

0.055

6.07

0.141

15.7

0.061

6.77

0.137

15.2

4.0

10

0.044

4.88

0.105

11.7

0.045

4.98

0.108

12.0

0.050

5.55

0.106

11.9

6.0

15

0.030

3.29

0.065

7.27

0.030

3.36

0.067

7.48

0.034

3.82

0.068

7.62

282

Table 5-5 Ratios of AUTODYN calculations to polynomial and UFC 3-340-02 predictions
for reflected scaled impulse
Z
1/3

m/kg

0.16
0.20
0.26
0.30
0.40
0.60
0.80
1.2
1.6
2.4
3.2
4.0
6.0
8.0

1/3

ft/lb

0.40
0.50
0.65
0.75
1.0
1.5
2.0
3.0
4.0
6.0
8.0
10
15
20

AUTODYN/polynomials
I s W 1/3
I r W 1/3
1.07
1.07
1.07
1.06
1.04
0.97
1.04
0.98
1.01
1.01
0.99
0.98
0.98
0.98

0.99
0.99
0.99
0.99
0.99
0.99
1.00
1.02
1.03
1.00
0.98
0.98
0.97
0.99

283

AUTODYN/UFC 3-340-02
I s W 1/3
I r W 1/3
0.87
0.81
0.82
0.87
1.03
1.12
0.94
0.79
0.82
0.88
0.89
0.88
0.86
0.85

1.08
1.03
0.97
0.99
0.89
0.9
0.94
1.03
1.07
1.05
1.01
0.98
0.95
0.95

5.8 Reflected Impulse as a Function of Angle of Incidence


Figure 5-22 presents results of AUTODYN calculations of reflected scaled impulse. Values are
provided in Table 5-6. The effect of Mach reflection on the calculated reflected scaled impulses
is not seen in these results. To examine why, selected reflected overpressure and scaled impulse
histories are presented in Figures 5-23, 5-24 and 5-25 for Z = 0.16, 0.40 and 1.2 m/kg1/3,
respectively, and = 30, 40, 45, 50 and 55. Figures 5-23a and 5-23b present reflected
overpressure histories calculated for Z = 0.16 m/kg1/3 using cell sizes of R/800 and R/400,
respectively, for 0.24 t 0.4 msec to judge the effect of mesh size on the calculated reflected
peak overpressure. These histories for all five angles are complex with multiple peaks due to the
effects of Mach reflection and the expanding detonation products. The reflected peak
overpressures for the two cell sizes are similar. The local increase observed in the reflection
coefficient curve for Z = 0.16 m/kg1/3 with = 50, as shown in Figure 5-18, is also identified in
these history plots. The fluctuations in the reflected overpressure histories are more pronounced
for the mesh with R/800 cells. Figure 5-23c presents the reflected overpressure histories for the
mesh with R/400 cells for the duration of the positive phase. The contribution to the reflected
peak impulse of the overpressure history in the time window of 0.24 to 0.40 msec, and around
the time of reflected peak overpressure, is small. Further, given that the duration of the positive
phase is greatest for = 30 and smallest for = 55, and that the reflected overpressure at 0.40
msec is greatest for = 30 and smallest for = 55, the effect of Mach reflection on reflected
peak overpressure is not carried over to reflected peak impulse. The reflected peak impulse
decreases gradually with an increasing angle of incidence, as seen in Figure 5-23d. Similar
results are seen for Z = 0.40 and 1.2 m/kg1/3 as presented in Figures 5-24 and 5-25, respectively,
except that a) the reflected peak overpressures for both scaled distances are associated with the
arrival of the incident or reflected shock front (depending on ), and b) there is no effect of the
expanding detonation products on the reflected overpressure histories for Z = 1.2 m/kg1/3, as
described in Section 5.6.

284

Figure 5-22 Reflected scaled impulse as a function of angle of incidence (Z in units of


m/kg1/3)

285

Table 5-6 Reflected scaled impulses as a function of angle of incidence, , and scaled
distance, Z, calculated using AUTODYN (units: MPa-ms/kg1/3; 1 MPa-ms/kg1/3 =
111 psi-ms/lb1/3)
Angle of incidence

Z
1/3

m/kg

1/3

ft/lb

10

20

30

40

50

60

70

80

90

0.16

0.40

9.83

9.52

8.78

7.63

5.98

4.15

2.62

1.25

0.275

0.20

0.50

6.37

6.20

5.72

4.970

4.02

2.97

1.81

0.892

0.182

0.26

0.65

3.86

3.76

3.48

3.031

2.49

1.89

1.23

0.624

0.138

0.30

0.75

2.96

2.89

2.67

2.34

1.91

1.48

0.996

0.538

0.131

0.40

1.0

1.80

1.76

1.64

1.45

1.20

0.968

0.690

0.393

0.166

0.140

0.60

1.5

1.01

0.989

0.924

0.829

0.702

0.584

0.466

0.301

0.180

0.173

0.80

2.0

0.709

0.697

0.656

0.596

0.509

0.424

0.348

0.263

0.224

0.183

1.2

3.0

0.454

0.446

0.424

0.386

0.336

0.285

0.234

0.186

0.177

0.118

1.6

4.0

0.329

0.324

0.310

0.287

0.259

0.225

0.184

0.139

0.105

0.092

2.4

6.0

0.199

0.198

0.192

0.182

0.169

0.155

0.135

0.097

0.078

0.069

3.2

8.0

0.138

0.137

0.134

0.130

0.124

0.116

0.103

0.078

0.065

0.054

4.0

10

0.105

0.104

0.103

0.100

0.097

0.092

0.084

0.064

0.058

0.044

6.0

15

0.065

0.065

0.065

0.064

0.062

0.061

0.057

0.044

0.043

0.030

8.0

20

0.048

0.048

0.047

0.047

0.046

0.046

0.044

0.034

0.030

0.022

286

(a) Reflected overpressure history for the


R/800 cell size; part range on time

(b) Reflected overpressure history for the


R/400 cell size; part range on time

(c) Reflected overpressure history for the R/400 cell size

(d) Reflected impulse history from (c)


Figure 5-23 Reflected overpressure and scaled impulse histories for cell sizes of R/400 and
R/800; = 30, 40, 45, 50 and 55; Z = 0.16 m/kg1/3
287

(a) Reflected overpressure history for the


R/800 cell size; part range on time

(b) Reflected overpressure history for the


R/400 cell size; part range on time

(c) Reflected overpressure history for the R/400 cell size

(d) Reflected impulse history from (c)


Figure 5-24 Reflected overpressure and scaled impulse histories for cell sizes of R/400 and
R/800; = 30, 40, 45, 50 and 55; Z = 0.40 m/kg1/3
288

(a) Reflected overpressure history for the


R/800 cell size; part range on time

(b) Reflected overpressure history for the


R/400 cell size; part range on time

(c) Reflected overpressure history for the R/400 cell size

(d) Reflected impulse history from (c)


Figure 5-25 Reflected overpressure and scaled impulse histories for cell sizes of R/400 and
R/800; = 30, 40, 45, 50 and 55; Z = 1.2 m/kg1/3
289

Figures 5-23, 5-24 and 5-25 document the differences in arrival time of the shock front for the
selected angles of incidence. The arrival times for the three scaled distances and = 30, 40
and 45 are similar. However, the arrival times for = 50 and 55 are smaller than those for
other angles. Consider Figure 5-26 that presents pressure fringes at t = 1.0 ms for Z = 0.4 m/kg1/3
with = 30, 40, 45, 50 and 55; R/800 cells are used for analysis. The incident shock fronts
have radiated to the same monitoring location (point 1) in all cases. For = 50 and 55, the
Mach stems form and arrive at the monitoring locations before the incident waves because the
Mach stem is a reinforced shock front that propagates faster than the incident wave. Similar
differences in the arrival times are observed in the reflected overpressures histories for Z = 0.16
and 1.2 m/kg1/3.
Figure 5-27 enables a comparison of the AUTODYN calculations and the empirical charts in
UFC 3-340-02 for reflected scaled impulse as a function of angle of incidence. For each scaled
distance, Z, considered in the AUTODYN analysis, three curves from UFC 3-340-02 are plotted
for values of Ps close to the AUTODYN-calculated value of peak incident overpressure. The
AUTODYN-calculated reflected scaled impulses for scaled distances between 0.16 and 8 m/kg1/3
( Ps = 19.7 and 0.016 MPa, respectively) are similar to the values reported in UFC 3-340-02. No
changes are recommended to the UFC 3-340-02 charts for reflected scaled impulse.

290

(a) = 30

(b) = 40

(c) = 45

(d) = 50

(e) = 55
Figure 5-26 Pressure fringes at t = 1.0 ms for Z = 0.4 m/kg1/3; R/800 cells

(a) Z = 0.16 m/kg1/3 (0.4 ft/lb1/3)

(b) Z = 0.20 m/kg1/3 (0.5 ft/lb1/3)

(c) Z = 0.26 m/kg1/3 (0.65 ft/lb1/3)

(d) Z = 0.30 m/kg1/3 (0.75 ft/lb1/3)

(e) Z = 0.40 m/kg1/3 (1.0 ft/lb1/3)

(f) Z = 0.60 m/kg1/3 (1.5 ft/lb1/3)

Figure 5-27 Reflected scaled impulse as a function of angle of incidence; Ps in units of MPa

292

(g) Z = 0.80 m/kg1/3 (2.0 ft/lb1/3)

(h) Z = 1.2 m/kg1/3 (3.0 ft/lb1/3)

(i) Z = 1.6 m/kg1/3 (4.0 ft/lb1/3)

(j) Z = 2.4 m/kg1/3 (6.0 ft/lb1/3)

(k) Z = 3.2 m/kg1/3 (8.0 ft/lb1/3)

(l) Z = 4.0 m/kg1/3 (10 ft/lb1/3)

Figure 5-27 Reflected scaled impulse as a function of angle of incidence; Ps in units of MPa
(cont.)
293

(m) Z = 6.0 m/kg1/3 (15 ft/lb1/3)

(n) Z = 8.0 m/kg1/3 (20 ft/lb1/3)

Figure 5-27 Reflected scaled impulse as a function of angle of incidence; Ps in units of MPa
(cont.)

294

SECTION 6
MODELING CONCRETE EROSION STRAIN FOR BLAST ANALYSIS OF
STRUCTURAL COMPONENTS
6.1 Introduction
Structural components subjected to extreme loadings such as blast effects, missile impact and
earthquakes often lose strength and stiffness due to damage and loss of material. Damage can be
modeled coarsely at the component level using macro-models (e.g., Mehanny and Deierlein.
2000, Lignos et al. 2011) or more finely using finite elements (e.g., Malvar and Simons 1996,
Murray 2007b). The latter approach is likely more accurate but computationally expensive.
Macro-models are best suited for earthquake analysis because all components in a structure
directly experience the load effect. Macro-models are generally based on regression analysis of
results of small-scale and large-scale tests and are not component specific. As such, macro
models are not particularly suitable for analysis of components (and structures) subjected to
more localized loadings such as detonations of 10s to low 1000s of kilograms of high explosives,
tornado-borne missile impact, and ballistics including bullets and shoulder-launched missiles.
For such loadings, detailed finite element (FE) models are preferred provided that robust
constitutive models are available. Component modeling using finite elements is the subject of
this study.
Blast and impact loadings can spall concrete from reinforced and pre-stressed concrete
components. Blast and missile impact loadings may be considered for the design of missioncritical and defense-related structures and form part of the design basis for safety-related nuclear
structures. High rates of straining are often generated in structural components due to the
impulsive loadings due to blast or impact. Any constitutive model that is used to predict the
response of structural components to blast and impact loadings should enable the analyst to
address the effects of loading rate.

295

Fragmentation of concrete may accompany a near-field detonation of high explosives. For


structural components such as columns that are susceptible to such loadings, the short duration of
loading (msec) and short period of oscillation of the component (10s of msec) often result in a
high rate of straining of the extreme fibers of components. Excessive strains will lead to material
failure and strain rate will affect the threshold strain (termed here as the erosion strain) at which
material is lost (or eroded) from the component. Erosion strain is usually tensile because strain
capacity is smaller in tension than compression.
Erosion algorithms for concrete have been used in numerical simulations of blast and penetration
events to eliminate material from a model (e.g., Schwer 2004, Luccioni et al. 2004, Teng et al.
2004, Xu and Lu 2006, Polanco-Loria et al. 2008, Islam et al. 2011). Although both the blast and
ballistic (penetration) loadings can generate large local deformations, high straining, and damage
or failure, they cannot be treated similarly because the loading conditions are quite different.
This study focuses on blast loadings although some techniques used for penetration studies of
concrete are described to aid in the determination of criteria for concrete erosion.
The effect of using alternate values of concrete erosion strain and alternate constitutive material
models (concrete and reinforcement) is investigated in this study by simulating the response of a
sample reinforced concrete (RC) column subjected to free-air TNT blast loadings using values of
erosion adopted by Xu and Lu (2006) and Luccioni et al. (2004). The results of these simulations
are compared with those performed using concrete erosion criterion established in this study. A
model of the sample RC column is built in a finite element (FE) code, LS-DYNA (LSTC 2013)
with pin-fixed boundary conditions to approximate a first story column founded in a raft
foundation. The TNT detonation is modeled using two codes: Air3D (Rose 2006) and
AUTODYN (ANSYS 2013a), to generate reflected pressure histories on the column. The
reflected pressure histories are then directly applied to the sample RC column in LS-DYNA.
To enable the simulation of the response of the column to air-blast loading, trial analyses are
performed to estimate the magnitude of the tensile stresses on the back face of the column
resulting from reflection of the incident compressive shock front. The results of the trial analyses
are described in Section 6.3. The effect of strain rate on material response and alternate material

296

models for concrete and steel reinforcement are described in Section 6.4, including the effects of
strain rate. Single three-dimensional (3D) solid (finite) element simulations of concrete are
described in Section 6.5 to evaluate the sensitivity of the values of concrete erosion strains to rate
of loading, element dimension, loading condition and concrete compressive strength. Various
dimensions of single elements are considered based on uniform h-refinement (Cook 2002). Each
element is loaded in pure tension because erosion criteria for concrete are typically formulated
using a maximum principal tensile strain (e.g., Teng et al. 2004, Xu and Lu 2006, Polanco-Loria
et al. 2008, Islam et al. 2011), as noted above. LS-DYNA and its implementation of the CSCM
model are used for the single element simulations. Force control and triangular loading functions
are used to represent blast loadings.

6.2 Criteria for Eroding Elements


Criteria for eroding concrete have been determined by judgment and use of empirical data,
generally calibrated from but not measured directly in experiments. Luccioni et al. (2004)
established an erosion strain as an incremental geometric strain of 0.075 for a concrete
compressive strength of 25 MPa by calibration of AUTODYN simulations of uniformly
supported concrete slabs subjected to blast loadings. However, incremental geometric strain is
not an appropriate erosion parameter for brittle materials such as concrete, as discussed later,
because it is a function of the effective strain rate, a scalar quantity, and the properties of
concrete vary greatly in compression and tension.
Erosion strain for concrete should be determined by parameters such as principal stress and
principal strain. Maximum principal tensile strain and minimum principal compressive strain
have been used for this purpose (e.g., Schwer 2004, Teng et al. 2004, Islam et al. 2011). Xu and
Lu (2006) numerically simulated spallation in reinforced concrete walls cast with 40 MPa
concrete, fixed on four sides, and subjected to blast loadings. They proposed a maximum tensile
principal strain of 0.010 as a concrete erosion strain on an empirical basis with consideration of
typical concrete strain at peak tensile stress, strain softening, strain rate and confinement effects.
Although both Luccioni et al. and Xu and Lu demonstrated good agreement between their

297

simulations and test data, neither provided a technical basis for determining concrete erosion
strain.
Relationships among material parameters such as erosion strain, strain capacity and strain rate
are discussed here first because an understanding of these relationships plays a pivotal role in
establishing erosion strain for concrete. An erosion strain for concrete should be proportional to
its tensile strain capacity, which in turn is a function of the stress state, concrete compressive and
tensile strengths, loading history, and rate of strain. The compressive and tensile strengths of
concrete increase with increasing strain rate (e.g., Dusenberry 2010). At low rates of strain, say

105 sec-1, the tensile strain capacity of concrete tends to decrease with an increase in
strength (Wight and MacGregor 2011). One might expect the tensile strain capacity to further
decrease with increasing strain rate. Fracture energy is introduced below to clarify the
relationship between the two.
Tensile strain capacity can be quantitatively measured by fracture energy, which is defined as the
area under the stress-strain or stress-displacement curves after the peak stress (Xu et al. 2006).
Fracture energy is used to model softening behavior of brittle materials and is proportional to
tensile strain capacity. Lambert and Ross (2000), Lu and Xu (2004), Schuler et al. (2006), Brara
and Klepaczko (2007), and Ruiz et al. (2009) have reported on the basis of test data that fracture
energy increases with increasing strain rate, which takes into account the increase in the tensile
strain capacity of concrete with strain rate. Grote et al. (2001), Ragueneau and Gatuingt (2003),
and Ruiz et al. (2009) showed through compressive and tensile tests that maximum compressive
and tensile stress, and maximum compressive and tensile strain increase with increasing strain
rate.
Most simulations using erosion algorithms have been performed using LS-DYNA and
AUTODYN. These programs provide multiple erosion parameters. Parameters based on von
Mises theory (e.g., effective stress, effective strain, effective plastic strain) are well suited for
isotropic materials like metals (e.g., steel, copper, iron) (e.g., Kurtaran et al. 2003, Fawaz et al.
2004). For anisotropic materials (e.g., concrete, orthotropic material), parameters such as

298

principal stress and principal strain have been employed (e.g., Schwer 2004, Teng et al. 2004,
Islam et al. 2011). LS-DYNA supports the following parameters to erode material:
Effective plastic strain, effp
Effective strain, eff
Effective stress, eff
Failure time
Maximum compressive pressure (positive in compression)
Maximum principal tensile strain
Maximum principal tensile stress
Maximum shear strain (= (1 3 ) / 2 )
Minimum principal compressive strain
Minimum tensile pressure (positive in compression)
Stress impulse
Threshold stress
Volumetric strain (= jj )
where 1 and 3 are the first and third principal strains, respectively, and jj are the components
of the strain tensor. The effective plastic strain, effp , effective strain, eff , and effective stress,

eff , are given by

effp

2 p p
ij ij dt
3

(6-1)

eff

2 dev dev
ij ij
3

(6-2)

eff

2 dev dev
ij ij
3

(6-3)

where ijp , ijdev and ijdev are the components of the plastic strain-rate tensor, deviatoric strain
tensor, and deviatoric stress tensor, respectively. When the failure time is exceeded by a problem
time, material is eroded. The threshold stress is specified with the Tuler-Butcher criterion (Tuler
and Butcher 1968) as
t

[max(0,
0

0 )]2 dt K f

(6-4)
299

where 1 is the maximum principal stress, 0 is a specified threshold stress ( 1 0 0 ), and


K f is the material constant for failure. The volumetric strain can be positive or negative

depending upon whether failure is in tension or compression. LS-DYNA allows the use of one or
more parameters to erode material in a given analysis (LSTC 2013).
AUTODYN enables material erosion using one of the following parameters:
Effective plastic strain, effp
Hydrostatic tensile pressure, pt
Incremental geometric strain, incr
Instantaneous geometric strain, inst
Material strain: maximum tensile failure strains ( 11ft , 22ft and 33ft ) and maximum shear strains
( 12fs , 23fs and 31fs )
Material stress: maximum tensile failure stresses ( 11ft , 22ft and 33ft ) and maximum shear
stresses ( 12fs , 23fs and 31fs )
Material stress/strain: coupling of the criteria of the material strain and material stress failures
Principal strain: maximum principal tensile strain and (1 3 ) / 2
Principal stress: maximum principal tensile stress and (1 3 ) / 2
Principal stress/stress: coupling of the criteria of the principal strain and principal stress
failures
Time step: minimum element time step
where the effective plastic strain, effp , is defined in the previous paragraph, ijft , ijfs , ijft and

ijfs are the components of maximum tensile failure strain tensor, maximum shear strain tensor,
maximum tensile failure stress tensor, and maximum shear stress tensor, respectively, 1 and 3
are defined previously, and 1 and 3 are the first and third principal stresses, respectively. The
incremental geometric strain, incr , is defined by

incr

2
ij ij dt
3

(6-5)

where ij are the components of the strain rate tensor. The incremental geometric strain, incr , is
a non-decreasing cumulative function with time and increases continuously for both compressive
and tensile stresses of elements, which can lead to erroneous material erosion in the elastic range,
300

especially for cyclic loading. ANSYS (2013a) introduces an instantaneous geometrical strain,

inst , in AUTODYN to overcome this shortcoming as

inst

2
(112 222 332 ) 5(11 22 22 33 3311 ) 3(122 232 312 )
3

(6-6)

where ij are the components of the strain tensor. Unlike the incremental geometric strain, incr ,
the instantaneous geometrical strain, inst , can decrease for cyclic loading of elements and
computes the equivalent strain in a similar manner to von Mises strain.
Material erosion can be also initiated by a limiting value of time step. A local time step is
computed for each element in a model as a fraction of the time required for the passage of the
dilatational wave across the minimum element dimension. The global time step for analysis of
the mathematical model is taken as the minimum value of the local time steps. If elements are
highly distorted, the minimum dimension of all distorted elements can be very small (indirectly
suggesting failure), resulting in a very small local time step and a significant increase in
computational effort. Elements can be eroded or deleted if the local time steps decrease to a userspecified minimum element time step to manage this effort.

6.3 Wave Passage Effects in Reinforced Concrete Columns


Wave passage effects through the cross section of reinforced concrete components are often
idealized using one-dimensional propagation of impulsive loadings in rods (e.g., Timoshenko
and Goodier 1970, Kolsky 1953). For a zero-damped rod, impulsively loaded at one end in
compression, the wave is reflected at the free end as a tensile wave with the same amplitude as
the incident compressive wave.
Impulsive loadings are imposed on the front face of a sample reinforced concrete column to
provide insight into the propagation of stress waves through the column and their subsequent
reflection from free surfaces: the back and front faces, respectively. The results of these
numerical analyses are contrasted with one-dimensional wave propagation solutions to
demonstrate that the classical one-dimensional solutions cannot predict stresses in threedimensional structures such as columns.

301

Consider the reinforced concrete column of Figure 6-1 that is 600 mm by 600 mm in plan and 4
meters in height. For the purpose of this calculation, the material is assumed to be linearly elastic
and the longitudinal and transverse reinforcement is ignored. (These assumptions are relaxed in
following sections.) The column is meshed with solid cubic elements with a 40 mm side length.
The Youngs modulus and Poissons ratio were assumed to be 28 GPa and 0.2, respectively.

(a) Plan view

(b) Elevation

Figure 6-1 Reinforced concrete column (units in meters)


To study wave propagation in this column, an impulsive load is applied per Figure 6-2. The
loaded area is at the midheight of the column and measured 1 meter by 600 mm (width). The
reflected peak overpressure is 120 MPa, which is similar to that computed later in Section 6.6.
The loading is imposed at t = 0.1 msec. The duration of the impulse is 0.02 msec, which is the
shortest possible based on the size of the element and the dilatational wave speed in concrete, c,
which is

E (1 )
3600 m/sec
(1 )(1 2 )

(6-7)

302

where E is Youngs modulus, is Poissons ratio and is the mass density of concrete (= 2400
kg/m3). For a mesh size of 40 mm, the duration of the pulse must be greater than 0.011 msec (=
40 mm/3600 mm/msec) for the loading to b1e resolved.

Figure 6-2 Rectangular pulse used for the elastic simulation of the sample RC column
Figure 6-3 is a cross section of the column at its midheight. The coordinate system is shown in
the figure; loading is in the x direction. Stress histories are generated for seven elements through
the depth, numbered 1 through 7 in the figure. Figure 6-4 presents histories in the x direction for
elements 1, 3 and 7, where tensile stress is positive. The travel time for the compression wave
through the column is 0.16 msec, with the wave arriving at the back face of the column at t =
0.26 msec and arriving back at the front face at t = 0.42 msec. Figure 6-5 presents stress histories
for elements 4, 5, 6 and 7. The amplitude in compression diminishes with distance traveled due
to wave scattering and the Poisson effect. The amplitude of the reflected wave (tensile) in
element 7 is greater (= 50 MPa) than the amplitude of the incident compressive wave (= 20 MPa).

303

Figure 6-3 Cross section at the mid-height of the sample RC column

Figure 6-4 Stress histories in the x-direction for elements No. 1, 3 and 7

Figure 6-5 Stress histories in the x-direction for elements No. 4, 5, 6 and 7

304

To study the influence of wave reflection for this column, the depth of the column is doubled to
1.2 m as shown in Figure 6-6 and the loading of Figure 6-2 is applied. Figure 6-7 presents the
stress histories in elements 1 through 7 in the x direction for both columns. The effects of
reflection from the free surface are evident in the 600-mm-deep column at t = 0.42 msec
(element 1), 0.41 msec (element 2), 0.33 msec (element 3), 0.29 msec (element 4), 0.28 msec
(element 5), 0.27 msec (element 6) and 0.26 msec (element 7). In panel (g), it can be seen that
the reflected tensile stress wave reduces the amplitude of the incident compressive stress wave,
which has an amplitude of approximately 60 MPa in panels (d) (element 4), (e) (element 5), and
(f) (element 6). Figure 6-8 presents the differences in the stress histories for the two columns in
elements 4, 5, 6 and 7; the differences (in tensile stress) are the result of wave reflection. The
amplitude of the tensile reflected wave of approximately 60 MPa is approximately equal to the
peak amplitude of the compressive wave in elements 5, 6 and 7 in the 1.2-m-deep column at t =
0.27 to 0.30 msec. In this instance, the amplitude of the incident compressive wave in element 7
of the 1.2-m-deep column (the free or rear surface of the 600-mm-deep column) is approximately
equal to the amplitude of the reflected tensile wave in the 600-mm-deep column.
This simple example has demonstrated that the extension of one-dimensional wave propagation
results to simple two- and three-dimensional shapes leads to erroneous results. The amplitude of
the compressive wave diminishes with distance traveled through the column; the reflection of the
compressive wave from the back face of the column destroys the wave field. The calculation of
stress histories is further complicated by a) the replacement of the simplified impulsive loading
on part of the front face of the column with a temporal and spatially varying pressure history
over the entire surface of the column, which is more representative of air-blast loading, b) the
substitution of elastic material by a nonlinear concrete model, and c) the addition of longitudinal
and transverse reinforcement to the column.

305

Figure 6-6 Cross section at the mid-height of the 1.2-m-deep column

(a) Element No. 1

(b) Element No. 2


Figure 6-7 Comparison of stress histories of elements No. 1 through 7 in the x-direction
between 0.6-m- and 1.2-m-deep columns

306

(c) Element No. 3

(d) Element No. 4

(e) Element No. 5


Figure 6-7 Comparison of stress histories of elements No. 1 through 7 in the x-direction
between 0.6-m- and 1.2-m-deep columns (cont.)

307

(f) Element No. 6

(g) Element No. 7


Figure 6-7 Comparison of stress histories of elements No. 1 through 7 in the x-direction
between 0.6-m- and 1.2-m-deep columns (cont.)

Figure 6-8 Differences in the stress histories of elements No. 4, 5, 6 and 7 in the x-direction
between the 0.6-m- and 1.2-m-deep columns

308

6.4 Material Models


6.4.1 Concrete
6.4.1.1 Strain Rate Effects
Strain-rate effects for concrete are different for tensile and compressive loadings. A Dynamic
Increase Factor (DIF) is often used to adjust the compressive and tensile strengths of concrete for
high strain rates, and forms a basis for blast resistant design per UFC 3-340-02 (DoD 2008).
Different values and equations for DIFs for concrete have been proposed (e.g., CEB 1993,
Malvar and Ross 1998, Hao and Zhou 2007). For compressive strength, the CEB
recommendation that is shown in Figure 6-9a is widely used (e.g., Dusenberry 2010, Hao and
Zhou 2007, Murray 2007, Brannon and Leelavanichkul 2009, Coughlin et al. 2010) and is
adopted for this study. For tensile strength, Hao and Zhou (2007) proposed DIFs based on
extensive experimental data (e.g., Malvar and Ross 1998, Schuler et al. 2006). Their formulation
for tensile strength is shown in Figure 6-9b together with the proposals of the CEB (1993) and
Malvar and Ross (1998). Note the significant difference between the three curves in panel (b),
which will inevitably impact the results of the simulations but cannot be addressed here.
6.4.1.2 Continuous Surface Cap Model (CSCM)
The Continuous Surface Cap Model (CSCM) (Murray 2007) is used to simulate the response of
concrete. It was developed for the United States Federal Highway Administration to model the
crashworthiness of concrete structures. The CSCM model includes a damage formulation, and
can accommodate strain-rate effects and erosion due to damage accumulation. Strain-rate effects
are modeled using a viscoplastic formulation. A general shape of the CSCM yield surface is
shown in Figure 6-10. The yield surface is composed of the shear (failure) surface and hardening
compaction surface (cap). The cap model smooths the intersection between the two surfaces to
eliminate the likelihood of numerical instabilities. The yield surface is continuous and symmetric
about the three principal stress axes and the hydrostatic axis (referred to as the pressure axis)
serves as its center line. The CSCM model is formulated using equations for the yield surface,

309

strain-rate effects, damage formulation and strain softening, which are sequentially described
below.

(a) DIF for compressive strength

(b) DIF for tensile strength


Figure 6-9 Dynamic Increase Factor for compressive and tensile strength of concrete

Figure 6-10 General shape of the CSCM yield surface (Murray 2007)
310

The yield surface of the CSCM model is formulated using three stress invariants (Chen 2007) as
follows,

f ( J1 , J 2 , J 3 , ) J 2 2 Ff2 Fc

(6-8)

where J1 is the first invariant of the stress tens or, J 2 is the second invariant of the deviatoric
stress tensor, Sij , J 3 is the third invariant of the deviatoric stress tensor, is the Rubin threeinvariant reduction factor, F f is a shear failure surface function, Fc is a hardening cap function,
and is a cap hardening parameter. The three stress invariants are defined by using the
deviatoric stress tensor, Sij , and pressure, p, as

J1 3 p

(6-9)

J2

1
Sij Sij
2

(6-10)

J3

1
Sij S jk Ski
3

(6-11)

The shear failure surface function, F f , is defined as

Ff ( J1 ) e J1 J1

(6-12)

where , , and are obtained by fitting the model surface to strength measurements from
triaxial compression (TXC) tests. The Rubin reduction factor is a scaling function that
determines the strength of concrete for any state of stress relative to the strength for the TXC
tests (Rubin 1991). The hardening cap function, Fc , is used to model plastic volume change
related to porosity. The function is expressed as
Fc ( J1 , k ) 1

[ J1 L(k )][ J1 L(k ) J1 L(k )]


2[ X (k ) L(k )]2

(6-13)

where L(k ) is

k
L(k )
k0

if

k k0

(6-14)

otherwise

311

where k0 is the value of J1 at the initial intersection of the cap and shear surfaces before the cap
moves. The function of L(k ) prevent the cap from retracting into its initial location at k0 . The
intersection of the cap with the J1 axis is at
X (k ) L(k ) RFf ( L(k ))

(6-15)

where R is the ellipticity ratio of the major to minor axes of the cap. The cap moves with change
in plastic volume. The motion of expansion and contraction is based on the hardening rule

vp W (1 e D ( X X
1

0 ) D2 ( X X 0 )

(6-16)

where vp is a plastic volume strain, W is a maximum plastic volume strain, D1 and D2 are
shape parameters, and X 0 is the initial location of the cap. The parameters of the plastic volume
strain, vp , are obtained by fitting pressure-volumetric strain curves to data.
As noted previously (and by many others), the compressive and tensile strength of concrete
increases with strain rate. The viscoplastic algorithm interpolates the elastic stress, ijT , and the
elasto-plastic stress, ijp , as follows,

ijvp (1 ) ijT ijp

(6-17)

t /
1 t /

(6-18)

where ijvp is the viscoplastic stress, which addresses strain-rate effects using a parameter, , as

0
N

(6-19)

where strain-rate effect parameters, 0 and N, are obtained by curve fitting (or calibration) to
strain-rate data and are fitted to the CEB recommendation for concrete compressive strength and
the Hao and Zhou recommendation for tensile strength.
Damage is addressed by considerations of strain softening and modulus reduction. The strain
softening indicates a decrease in strength after peak strength is reached. Modulus reduction is a
decrease in the unloading or loading slopes. The stress, ijd , is

312

ijd (1 d ) ijvp

(6-20)

where d is a damage parameter with a value between 0 and 1, and ijvp is the viscoplastic stress
tensor without damage. Damage accumulation is either ductile or brittle. Ductile damage
accumulates when the pressure is compressive and depends on the total strain components, ij ,
as
c

1
ij ij
2

(6-21)

where c is an energy-type term, and ij are the elasto-plastic stresses. Ductile damage initiates
when c exceeds a ductile damage threshold, r0c , which is associated with the onset of softening.
Brittle damage accumulates when the pressure is tensile and depends on the maximum principal
strain, max , as
2
t E max

(6-22)

where t is an energy-type term, and E is the initial elastic modulus of concrete. Brittle damage
initiates when t exceeds the brittle damage threshold, r0b , that shifts as a function of strain-rate
as follows,

E
r0 1
r E
s

rs

(6-23)

where r0 is the shifted damage threshold, rs is the damage threshold before applying
viscoplasticity, and is the strain-rate effect parameter introduced in Equation 6-19. The
damage threshold grows with the increase in strain rate. As damage accumulates, the damage
parameter, d, increases from a minimum value of zero toward a maximum value of 1.0. The
damage parameters for compression and tension, d ( b ) and d ( t ) , respectively, are computed
during the softening phase, namely,
d ( c )

d max
1 B

1
A( c r0 c )

B 1 Be

(6-24)

d ( t )

0.999
1 D
1
D 1 DeC ( t r0 t )

(6-25)

313

where the parameters, A, B, C and D, are obtained by curve-fitting to stress-strain data, d max is a
maximum damage level, defined as

3J 2
d max J1

1.5

if

3J 2
1
J1

(6-26)

if otherwise

The maximum damage varies with strain-rate effects as (Murray 2007b)

d max d max max1.0,

1.5

r E
s

(6-27)

Elements are eliminated when the damage parameter, d, exceeds 0.99.


Finite element analyses of models of materials with softening formulations such as concrete have
been known not to converge due to strain localization (e.g., Baant 1986, Belytschko et al. 1986,
Baant and Chang 1987, Baant and Pijaudier-Cabot 1988). Consider a cube of concrete
subjected to normal and shearing tractions, which is modeled with fine and coarse meshes.
Damage accumulation will be greater in strain-localized elements of the finely meshed model,
and the elements will then be eroded from this model first, which is a result of fracture energy
being smaller in the smaller elements. Sandler and Wright (1984) asserted that the strainsoftening models provide incorrect solutions because small differences in loading can lead to
large changes in response. To remedy this situation, nonlocal formulations have been applied
(e.g., Baant and Chang 1987, Baant and Pijaudier-Cabot 1988). The nonlocal formulation
eliminates element-to-element variations in softening behavior by controlling fracture energy.
Murray (2007) showed through simulations of tensile tests of concrete that converged solutions
can be attained if the fracture energy is independent of mesh size. The fracture energy, G f , is
expressed as

G f (1 d ) f dx

(6-28)

x0

where x0 is a displacement at peak strength, f , and x is displacement. The fracture energy is


computed using the damage parameter, d, as given in Equations 6-24 and 6-25, and element

314

dimensions. The CSCM model allows users to specify values of fracture energy for uniaxial
tensile stress, G ft , pure shear stress, G fs , and uniaxial compressive stress, G fc . The fracture
energies for uniaxial tensile stress and pure shear stress correspond to the flexural and brittle
failure modes, respectively. Strain-rate effects on the three fracture energies are formulated in the
same way, namely,

E
G vp

f Gf
1 s
r E

repow

(6-29)

where G vp
f is the viscoplastic fracture energy, repow is a constant and other parameters were
defined previously.
6.4.2 Steel Reinforcement
Johnson and Cook (1983) proposed a constitutive model for metals subjected to high strain rates
and high temperature. It is used in this study to model reinforcement in the sample RC column of
Section 6.6. The basic form of the Johnson and Cook (JC) model is

y ( A B pn )(1 C ln

p
)(1 [T *]m )
0

(6-30)

where y is the yield stress, A is the yield stress per a reference strain rate, 0 , B and n are
material constants that represent the effects of strain hardening, p is the equivalent plastic strain,
C and m are experimentally determined constants, p is the equivalent plastic strain rate, and T*
is a normalized temperature, given by
T*

T Tr

(6-31)

Tm Tr

where T is the temperature, Tr is the room temperature, and Tm is the melting point. The values
of the constants for various metals in Equation 6-30 have been reported by Johnson and Cook
(1983), Zukas (1990), and Meyers et al. (1992), among others. Thermal effects can generally be
neglected for blast loadings on steel elements because the speed of heat conduction through the
steel is much slower than the speed of the shock front (Ballantyne et al. 2009).

315

The Johnson-Cook (JC) constitutive model has been used widely for computing the response of
metals deforming at high rates of strain. The model is available in LS-DYNA. Unfortunately, the
values of the parameters of the JC model for typical grades of reinforcement (Grades 60 and 75,
with yield stresses of 420 and 520 MPa, respectively) have not been published, requiring analysts
and researchers to adopt published values determined for similar metals (e.g., Johnson and Cook
1983, Gray et al. 1994). Danielson et al. (2008) used the JC model for Grade 40 steel
reinforcement (yield stress of 280 MPa, 40 ksi) but provided no information on the parameters of
the model. Zhou et al. (2008) used the JC model for AISI 4340 steel to model steel reinforcement
although the yield stress for AISI 4340 steel of 792 MPa is much higher than those of all grades
of ASTM A615 reinforcement, which is routinely used for commercial reinforced concrete
construction. Brvik et al. (2001) used values of the parameters for Weldox 460E, a high
strength structural steel, to model Grade 60 rebar. Weldox 460E has minimum specified values
for yield stress, ultimate stress and minimum elongation of 490 MPa, 580 MPa and 22%,
respectively, which are reasonably close to the corresponding values of 420 MPa, 620 MPa and 7
to 9%, for ASTM A615 Grade 60 reinforcement. Accordingly, the JC parameters are used
herein for Weldox 460E to model Grade 60 steel reinforcement but 420 MPa is substituted for
the yield strength parameter, A. The values are tabulated in Table 6-1. The calculated stressstrain relationship for Grade 60 reinforcement at a reference strain rate equal to 510-4 s-1 is
shown in Figure 6-11.
Table 6-1 Assumed model parameters for Grade 6d0 reinforcement (adopted from Brvik
et al. 2001)
Parameter
Youngs modulus, E (GPa)
Poisson ratio, v
Density, (kg/m3)
Yield strength, A (MPa)
Strain hardening parameter, B (MPa)
Strain hardening parameter, n
Constant, C
Constant, m
Reference strain rate, 0 (s-1)

316

Value
200
0.33
7850
420
383
0.45
0.0114
0.94
510-4

Figure 6-11 Stress-strain relationship for Grade 60 rebar at a strain rate of 510-4 s-1
The JC model requires an Equation of State (EOS) as the hydrodynamic models compute only
deviatoric stresses. The EOS provides a pressure-volume relationship for materials subjected to
compression or tension.
The Gruneisen EOS is widely used in numerical simulations for metals to compute pressure
(Meyers et al. 1992). The pressure, p, for materials in the compressed state is computed as

0C 2 1 (1

0
2

a 2

S3
1 ( S1 1) S 2

2
1
( 1)

( 0 a ) E

(6-32)

and for expanded materials as


p 0C 2 ( 0 a ) E

(6-33)

where o is the initial density, C is the sound speed in the material, S1 , S 2 and S3 are the
coefficients of the slopes of a shock speed versus particle speed curve, 0 is a Gruneisen
coefficient, a is the first order volume correction to 0 , E is the internal energy per initial unit
volume, and is given by

1
0

(6-34)

317

where is the current density. The parameters of the Gruneisen EOS in Equation 6-32 and 6-33
are obtained by Tan et al. (2009), as shown in Table 6-2.
Table 6-2 Gruneisen EOS parameters for steel (Tan et al. 2009)

o (Kg/m3)

c (m/s)

S1

S2

S3

7896

4569

1.49

2.17

0.5

Johnson and Cook expanded their constitutive strength model (1983) to a damage model (1985)
as follows,
D

(6-35)

where D is a damage index, p is the increment of accumulated effective plastic strain, and f
is accumulated plastic strain to failure. The damage index D is zero for a virgin material and
failure occurs when D reaches 1.0. The accumulated plastic strain at failure, f , is

f D1 D2 exp D3

m
eff

*
1 D4 ln p 1 D5T *

(6-36)

where D1 through D5 are material constants, m is a hydrostatic stress, eff is an effective


stress, *p is a dimensionless strain rate, given by p / 0 , and 0 is a reference strain rate.
Equation 6-36 accounts for the effects of temperature, strain rate, and stress triaxiality. In FE
simulations, elements are removed when the corresponding damage index equals 1.0. The
parameters of the JC damage model for Weldox 460 E steel (Brvik et al. 2001) are presented in
Table 6-3. These values are adopted for the Grade 60 steel reinforcement modeled in the sample
RC column of Section 6.6.
Table 6-3 Johnson and Cook damage model parameters for Weldox 460 E steel (Brvik et
al. 2001)
D1

D2

D3

D4

D5

0.0705

1.732

0.54

0.015

318

6.5 Single Element Simulations


Single element simulations are performed in LS-DYNA using the CSCM model of Section
6.4.1.2 to evaluate concrete erosion strain with respect to strain rate, element size, compressive
strength, and loading condition. The values of the parameters of the CSCM model for
compressive strengths of 35.5 and 50 MPa 9 are listed in Appendix E. Values for strain-rate
effects in compression and tension are obtained based on the CEB and Hao and Zhou
recommendations, respectively.
The dimensions of the single elements are first considered to be 202020, 404040 and
808080 mm based on uniform h-refinement (Cook 2002), noting that the size of the smallest
element is on the order of a piece of aggregate. Elements sizes of 303030 and 606060 mm
are also considered for completeness. Figure 6-12 enables a comparison of compressive stressstrain curves of the CSCM model for different element sizes and the concrete model reported by
Popovics (1970), all for a compressive strength of 35.5 MPa. The Popovics curve lays between
the CSCM curves for element dimensions of 202020 and 404040 mm, which is consistent
with Murrays description of a reasonable element size (19 to 38 mm).

Figure 6-12 Compressive stressstrain curves for concrete models of the CSCM and by
Popovics (1970) for the compressive strength of 35.5 MPa

Normal and high strength concrete, respectively. Murray (2007) reported data for f c = 35 MPa, and

some of this data used below.

319

Strain capacity decreases with increasing element size in the CSCM model due to the use of
constant fracture energy. Since the fracture energy is independent of element size, larger
elements have less strain capacity per unit volume than smaller elements, as supported by Figure
6-12.
The forcing functions for the single element simulations are defined by traction rates between 50
and 400 MPa/msec, based on the results of simulations that are described in Section 6.6. The
peak tensile tractions of the loadings are set equal to 10, 20 and 40 MPa based on the product of
a) quasi-static tensile strength of about 10 percent of the compressive strength, and b) an increase
associated with strain rate. Erosion did not occur for the peak tensile traction of 5 MPa. The
duration of each loading is assumed to be 0.5 msec, which is appropriate for the weapon size and
stand-off distance considered later. Figure 6-13 shows the loading conditions of tensile traction
histories for a peak tensile traction of 10 MPa. The loadings for the peak tensile tractions of 20
and 40 MPa are generated in a similar manner.

Figure 6-13 Loading conditions used in single element simulations when the peak tensile
load = 10 MPa
Alternate restraint conditions (pinned or fixed) on one boundary are considered for the single
element simulations to understand their impact on the calculated values of the erosion strain. The
other end of the element is subjected to tensile loading. The values of erosion strain are not
significantly affected by the choice of boundary condition. Results are presented in Table 6-4.

320

The remaining single element simulations are performed assuming one end of each element is
fixed.
Table 6-4 Influence of boundary conditions in the single element simulations for a
compressive strength of 35.5 MPa at 200 MPa/msec
Element size

202020 mm

404040 mm

808080 mm

Peak tensile
traction

Erosion strain
Error (%)

Simply
supported

Fixed

10 MPa

0.0080

0.0074

20 MPa

0.0095

0.0093

10 MPa

0.0037

0.0034

20 MPa

0.014

0.013

10 MPa

0.0032

0.0030

20 MPa

0.016

0.016

Erosion is assumed when the damage parameter for tension (Equation 6-25) equals or exceeds
0.99. Traction rate is used instead of strain rate to observe concrete erosion strains in this study
for two reasons: 1) strain rate increases with traction rate and is proportional to velocity, which
changes continuously with time for blast-type loadings (see Figure 6-13), and 2) strain rate
increases rapidly during the softening phase of the concrete. Given that the purpose of the study
is to evaluate concrete erosion strain for blast loadings rather than for a particular strain rate,
parameters that best characterize the loading conditions such as traction rate and peak tensile
traction are used to present the results of the single element simulations. Three single-degree-offreedom (SDOF) calculations are performed to validate the LS-DYNA calculations of strain rate;
results are presented in Appendix F.
The maximum time step for explicit analysis is established by the Courant condition: the length
of the smallest element divided by the sound speed in the material (= 3600 m/sec for concrete).
The effect of the time step on the results is investigated by performing analysis using time steps
less than the maximum value: 0.7, 0.5, 0.3 and 0.01 times the maximum value. Analysis is
performed for the 404040 mm element subjected to the blast-type loadings of Figure 6-13
with peak tensile tractions of 10 and 20 MPa. Results are presented in Figure 6-14. Default in
321

the legend corresponds to the maximum time step. The five erosion strain versus traction rate
relationships are similar except for the curve of the maximum time step. A factor of 0.5 on the
default time step is used hereafter because the results are essentially unchanged for factors less
than 0.5.

(a) Peak tensile traction = 10 MPa

(b) Peak tensile traction = 20 MPa


Figure 6-14 Time step convergence analysis for an element size of 404040 mm
Figures 6-15 to 6-18 present erosion strain versus traction rate for different loading conditions
and element sizes. Figure 6-15 illustrates the relationships between concrete erosion strain,
traction rate and peak tensile tractions of 10, 20 and 40 MPa for each size of the single element:
the assumed compressive strength is 35.5 MPa. Figure 6-16 presents similar data for a
compressive strength of 50 MPa.
The erosion strains are similar for a given size of element across traction rates between 50 and
400 MPa/msec, both compressive strengths and all three peak tensile tractions. Figures 6-17 and

322

6-18 recast the data of Figures 6-15 and 6-16, respectively, to observe the effects of element size
on erosion strain. The smaller the size of the element, the greater the erosion strain because the
fracture energy is independent of element size.

`
(a) 20 mm element

(b) 30 mm element

(c) 40 mm element
Figure 6-15 Erosion strain versus traction rate in single element simulations for different
peak tensile tractions, f c = 35.5 MPa

323

(d) 60 mm element

(e) 80 mm element
Figure 6-15 Erosion strain versus traction rate in single element simulations for different
peak tensile tractions, f c = 35.5 MPa (cont.)

324

(a) 20 mm element

(b) 30 mm element

(c) 40 mm element
Figure 6-16 Erosion strain versus traction rate in single element simulations for different
peak tensile tractions, f c = 50 MPa

325

(d) 60 mm element

(e) 80 mm element
Figure 6-16 Erosion strain versus traction rate in single element simulations for different
peak tensile tractions, f c = 50 MPa (cont.)

326

(a) Peak tensile traction = 10 MPa

(b) Peak tensile traction = 20 MPa

(c) Peak tensile traction = 40 MPa


Figure 6-17 Erosion strain versus traction rate in single element simulations for different
element sizes, f c = 35.5 MPa

327

(a) Peak tensile traction = 10 MPa

(b) Peak tensile traction = 20 MPa

(c) Peak tensile traction = 40 MPa


Figure 6-18 Erosion strain versus traction rate in single element simulations for different
element sizes, f c = 50 MPa

328

6.6 Blast Analysis of a Sample Reinforced Concrete Column


6.6.1 Introduction
A sample RC column is subjected to air-blast loading to illustrate the differences in damage
predictions that arise due to the use of alternate models for erosion strain. The column has a
square cross section with plan dimensions of 0.61 m 0.61 m (22 ft) and a height of 4 m (= 13
ft). The vertical reinforcement ratio is 2 percent (12 #9 rebar of ASTM A615 Grade 60
uniformly distributed on all faces of the column; see ASTM 2003). The transverse reinforcement
is #5 rebar perimeter seismic (or closed) hoops per ACI 318-11 (ACI 2011) with a vertical
spacing of 10 cm (= 4 in). The 20 kg (= 44.1 lb) TNT explosive charge is spherical and located at
the half height of the column at a stand-off distance of 0.5 m (= 20 in) as shown in Figure 6-19.
The scaled distance, Z, is 0.19 m/kg1/3: a near-field detonation for which damage is expected.

(a) Monitoring locations of reflected pressures

(b) Application of recorded reflected pressures

Figure 6-19 Blast loading calculations (units: m)

329

The analysis of the column is performed in LS-DYNA (LSTC 2013). Concrete and
reinforcement are modeled in LS-DYNA using MAT_CSCM and MAT_JOHNSON_COOK,
respectively. Perfect bond between concrete and reinforcement is assumed. The air-blast pressure
loadings on the front face of the column are computed using Air3D (Rose 2006), and
AUTODYN (ANSYS 2013a) and then applied to the model of the sample RC column. A mesh
convergence study is performed to identify reasonable element sizes for the simulations. Erosion
strains of concrete are calculated for the element sizes established in the mesh convergence study
using the single element simulations described in the previous section.
The blast analysis of the sample RC column is performed using the concrete erosion strains
calculated in this study. Analysis is also performed using a) the erosion strains of 0.075 and 0.01
used by Luccioni et al. (2004) and Xu and Lu (2006), respectively, and b) the damage functions
of the CSCM model that do not require the user to specify an erosion strain. Simulations are
performed to ensure transportability to other concrete models of the erosion strains computed
using the CSCM model. The influence of concrete compressive strength, confinement using
transverse reinforcement, and axial pressure loading on damage to the column are also addressed
through simulation.
6.6.2 Blast Loading
The histories of the reflected overpressures are computed along the height of the sample RC
column using Air3D and AUTODYN. One-dimensional analysis is performed for the explosion
in free air before the shock front reaches a reflecting surface (the target or the ground in this
instance) and the results of the one-dimensional analysis are then mapped into a threedimensional domain.
For the purpose of the simulations, pressure histories or time series are monitored in Air3D and
AUTODYN on the centerline of the sample RC column at the five locations shown in Figure 619(a). The pressure histories calculated at the monitoring locations 1, 2, 3, 4 and 5 are applied to
the column modeled in LS-DYNA as loadings 1, 2, 3, 4 and 5, respectively, as shown in Figure
6-19(b). The reflected peak pressure at monitoring location 3 is approximately 10 times greater
than at monitoring locations 2 and 4. The reflected pressure histories obtained from Air3D and
330

AUTODYN are similar as presented in Figure 6-20. The reflected peak pressures at monitoring
location 1 are greater than at monitoring location 5 due to the reflection of the shock front at the
ground surface. The AUTODYN results are used for the simulations because of concerns
regarding the accuracy of Air3D for near-field detonations (Anderson 2003) although the
differences between the AUTODYN and Air3D results are small.
The use of these pressure histories for analyses will overestimate the loading on the column
because clearing will reduce the reflected pressure at the edges of the column and pressure
loading on the back face of the column will reduce the net loading. Axial load will accumulate in
the column due to pressure loading on the underside of the floor system that the column is
supporting but this is not considered here.
6.6.3 Mesh Convergence Study
A mesh convergence study is performed to determine a reasonable balance between solution
accuracy and computation expense. Five finite element meshes (20 20 20, 30 30 30,
404040, 606060 and 808080 mm) of the sample RC column are prepared and
analyzed using the AUTODYN pressure histories. Erosion is not implemented for these
calculations.
The points of reference for the mesh convergence study are the horizontal displacement history
at the mid-height of the column and the horizontal reaction history at the bottom of the column.
Results are presented in Figure 6-21. Although the horizontal reaction histories are similar for all
mesh sizes, the displacement history for the 808080 mm mesh is somewhat different from
those for the 202020, 303030, 404040 and 606060 mm meshes. Accordingly, the
four smaller mesh sizes are used for the simulations that included erosion.
The mesh convergence study provides an initial estimate of the response of the sample column.
The maximum displacement of the column at its mid-height is on the order of 20 mm, or
span/200, which is unlikely to produce significant flexural damage. The maximum reaction is
3000 kN, which greatly exceeds the ACI limit of Vn 0.83 fc' bwd =1840 kN ( 10 fc' bw d in US

331

units). Although strain rate effects may substantially increase shear resistance above this ACI
limit, the magnitude of the peak shearing force indicates that damage near the support is likely.

(a) Monitoring location 1

(b) Monitoring location 2

(c) Monitoring location 3

(d) Monitoring location 4

(e) Monitoring location 5


Figure 6-20 Reflected pressure histories on the sample reinforced concrete column
332

(a) Horizontal displacement histories at mid-height

(b) Horizontal reaction histories at bottom


Figure 6-21 Mesh convergence study of the sample RC column
6.6.4 Values of Concrete Erosion Strain
If discrete values of erosion strain are to be used to predict damage in a reinforced concrete
column, the expected tensile pressures (tractions) and traction rates in the regions of the expected
damage must be established. For the sample column, the greatest damage in terms of eroded
material is expected on the rear face of the column, near its mid-height. To guide the calculation
of these tractions and traction rates, two analyses of the column are performed using different
material models and assuming a concrete compressive strength of 35.5 MPa: 1) elastic finite
elements with the small-strain values of modulus and Poissons ratio, and 2) a composite of
CSCM and elastic elements. Figure 6-22 shows the second model for the 202020 and
404040 mm meshes, where the 30 solid filled elements are elastic and the remainder are
CSCM. The small number of embedded elastic elements will not affect the global response of the
column. Erosion is not considered. Figure 6-23 presents the maximum principal stress histories
of an elastic element near the rear face and mid-height of the column for the two analyses. Model
1 is denoted Elastic; model 2, CSCM embeds 30 elastic elements in the mesh. The results of

333

the analysis of the Elastic models grossly overestimate the expected peak tensile traction and
traction rates and so the computationally more expensive CSCM model must be used.

Side elevation

Rear elevation

Side elevation

(a) 20 mm mesh

Rear elevation

(b) 40 mm mesh

Figure 6-22 FE models of the sample RC column to estimate traction rates and peak tensile
stresses (tractions) for 30 elements near or on the rear face of the column at its
mid-height

Figure 6-23 Maximum principal tensile stress history of an elastic element in models 1 and
2, 40 mm mesh
The maximum (tensile) principal stress histories for the 30 elastic elements in model 2 are
presented in Figure 6-24. The peak tensile stresses (tractions) range between 15 and 20 MPa and
the tensile stress (traction) rates are between 150 and 200 MPa/msec. A number of loading pulses
are evident in these stress histories, which are associated with the arrival of the first incident and
subsequent reflected shock waves.

334

(a) 202020 mm mesh

(b) 404040 mm mesh

Figure 6-24 Maximum principal tensile stress histories for 30 elements near or on the rear
face of the column at its mid-height; model 2
Values of erosion strain can be estimated for concrete compressive strengths of 35.5 and 50 MPa
from Figures 6-15, 6-16, 6-17 and 6-18. From Figure 6-15, for element sizes of 202020,
303030, 404040 and 606060 mm, 35.5 MPa concrete, peak tensile tractions between
15 and 20 MPa, and loading rates between 150 and 200 MPa/msec, the erosion strains are
approximately 0.01, 0.008, 0.006 and 0.005, respectively. These four values of erosion strains
are used in the simulations presented in the following section.
6.6.5 Simulation Results
Simulations of the blast response of the sample RC column are performed using 202020,
303030, 404040, and 606060 mm meshes of the column. Calculations are made for
the four mesh-dependent values of erosion strain identified above, the damage-based erosion
algorithm of Section 6.4.1.2, and the mesh-independent values of 0.01 (Xu and Lu, 2006) and
0.075 (Luccioni et al. 2004).
Figures 6-25 through 6-28 present displacement histories at the mid-height of the column on the
loading face for the 20 20 20, 30 30 30, 40 40 40, and 60 60 60 mm meshes,
respectively. Horizontal reaction histories at the bottom of the column are presented in Figures
6-29 through 6-32. The displacement and reaction are positive in the direction of the loading.
The displacement histories are essentially independent of mesh size for a given value of erosion
strain or if the damage-based algorithm is used. The peak displacement and time to peak
displacement calculated with considerations of erosion are similar to those of Figure 6-21. The

335

implementation of erosion results in substantially greater calculated values of residual


displacements, and, as expected, the smaller the value of the erosion strain, the greater the
residual displacement. The displacement history computed using an erosion strain of 0.075 is
most similar to that of Figure 6-21, where erosion was not considered.

Figure 6-25 Displacement histories at the mid-height of the column for the 202020 mm
mesh

Figure 6-26 Displacement histories at the mid-height of the column for the 303030 mm
mesh

336

Figure 6-27 Displacement histories at the mid-height of the column for the 404040 mm
mesh

Figure 6-28 Displacement histories at the mid-height of the column for the 606060 mm
mesh

Figure 6-29 Reaction histories at the bottom of the column for the 202020 mm mesh

337

Figure 6-30 Reaction histories at the bottom of the column for the 303030 mm mesh

Figure 6-31 Reaction histories at the bottom of the column for the 404040 mm mesh

Figure 6-32 Reaction histories at the bottom of the column for the 606060 mm mesh

338

The peak reactions are similar for all meshes and all simulations of erosion. The times to peak
reactions are smaller than the corresponding times to peak displacement. The reaction is negative
at approximately 0.75 msec, which is counter-intuitive. The negative reaction at this time instant
is due to wave propagation, up and down the column, where the wave fields are complex due to
reflection of compressive and tensile waves from the four free surfaces of the column. Figures
6-33 through 6-36 present internal energy histories for the 202020, 303030, 404040,
and 606060 mm meshes, respectively. The internal energy (or deformation energy) is the sum
of energies computed incrementally based on six components of stress and strain for all elements
in the mesh. Internal energy is not recorded for eroded elements.

Figure 6-33 Internal energy histories for the 202020 mm mesh

Figure 6-34 Internal energy histories for the 303030 mm mesh

339

Figure 6-35 Internal energy histories for the 404040 mm mesh

Figure 6-36 Internal energy histories for the 606060 mm mesh


The internal energies for the two simulations using an erosion strain of 0.075 are considerably
greater than for the other simulations because very few elements are eroded from the model as
seen in Figures 6-37(a), 6-38(a), 6-39(a) and 6-40(a). For the 202020 mm mesh, the internal
energy histories associated with the mesh-dependent erosion strain of 0.01 and the damage-based
erosion algorithm are very similar, with peak values of approximately 60 kJ. Similar trends are
seen for other meshes. The internal energies do not return to zero after loading because of the
permanent (residual) deformation of the column, as shown in Figures 6-25 through 6-28.
Figures 6-37 through 6-40 describe the damage to the sample column using side elevations and
cross sections at the mid-height of the column for the simulations using the 20 20 20,
303030, 404040, and 606060 mm meshes, respectively. Information is presented at a

340

time equal to 25 msec. Damage is acute for all meshes and simulations except those using an
erosion strain of 0.075. Consistent with the internal energy histories of Figures 6-33 through 6-36,
the extent of the damage to the column is similar for the simulations using the mesh-independent
erosion strains and the damage-based erosion algorithm: panels (b) and (c) of Figure 6-37 and
panels (c) and (d) of Figures 6-38, 6-39 and 6-40. Table 6-5 summarizes the percentage of the
cross section lost at the mid-height of the column for each of the simulations. Axial load was not
applied to the sample column but would have exacerbated the calculated damage. The effect of
axial load on the predicted damage is presented in the following section.

Table 6-5 Percentage erosion of the cross section at the mid-height of the column

Erosion
strain

202020 mm

303030 mm

404040 mm

606060 mm

0.075

7%

5%

7%

0%

0.01

77 %

55 %

46 %

44 %

0.008

Not calculated10

69

Not calculated

Not calculated

0.006

Not calculated

Not calculated

64 %

Not calculated

0.005

Not calculated

Not calculated

Not calculated

65

73 %

75 %

70 %

69

Damage function

341

Side elevation

Cross section at mid-height


(a) Erosion strain=0.075

Side elevation

Cross section at mid-height


(b) Erosion strain=0.01

Side elevation

Cross section at mid-height


(c) Damage function

Figure 6-37 Simulation results for the 202020 mm mesh at time = 25 msec

342

Side elevation

Cross section at mid-height


(a) Erosion strain=0.075

Side elevation

Cross section at mid-height


(b) Erosion strain = 0.01

Side elevation

Cross section at mid-height


(c) Erosion strain = 0.008

Side elevation

Cross section at mid-height


(d) Damage function

Figure 6-38 Simulation results for the 303030 mm mesh at time = 25 msec

343

Side elevation

Cross section at mid-height


(a) Erosion strain=0.075

Side elevation

Cross section at mid-height


(b) Erosion strain = 0.01

Side elevation

Cross section at mid-height


(c) Erosion strain = 0.006

Side elevation

Cross section at mid-height


(d) Damage function

Figure 6-39 Simulation results for the 404040 mm mesh at time = 25 msec

344

Side elevation

Cross section at mid-height


(a) Erosion strain=0.075

Side elevation

Cross section at mid-height


(b) Erosion strain = 0.01

Side elevation

Cross section at mid-height


(c) Erosion strain = 0.005

Side elevation

Cross section at mid-height


(d) Damage function

Figure 6-40 Simulation results for the 606060 mm mesh at time = 25 msec

345

6.6.6 Alternate Concrete Material Models, Influence of Concrete Compressive Strength,


Confinement, and Axial Pressure Loading
6.6.6.1 Introduction
The sample RC column is further analyzed to understand 1) the utility of the discrete values of
erosion strain when used with alternate concrete material models, 2) the influence of concrete
compressive strength on the volume of eroded material, 3) the effect of confinement in the form
of transverse reinforcement on the volume of eroded material, and 4) the effect of axial loading
on the column on residual lateral displacements and damage. Mesh sizes of 202020 mm,
303030 mm and 404040 mm are used for these analyses. The mesh-independent values of
0.01 (Xu and Lu, 2006) and 0.075 (Luccioni et al. 2004) are not considered.
6.6.6.2 Alternate Concrete Material Models
The CONCRETE_DAMAGE_REL3 model (Malvar and Simons 1996) is used to judge the
utility of the erosions strains proposed previously to predict damage using material models other
than the CSCM. A concrete compressive strength of 35.5 MPa is used for the analysis.
Figure 6-41 shows the maximum principal stress histories for the 30 elastic elements near or on
the rear face of the column, near its mid-height, which are estimated for the alternate material
model. The peak tensile stresses and stress rates shown in this figure are very similar to those
calculated for the CSCM model and shown in Figure 6-24. Accordingly, the values of erosion
strain used for the analysis of the CSCM model are adopted for analysis of the
CONCRETE_DAMAGE_REL3 model. Figure 6-42 enables a comparison of the internal energy
histories for the CSCM model using calculated erosion strains and the damage-based erosion
algorithm, and the CONCRETE_DAMAGE_REL3 model using calculated erosion strains, for
the 202020 mm, 303030 mm and 404040 mm meshes. The results are virtually
identical for all three meshes and the two concrete models. Descriptions of damage are presented
in Figure 6-43. The eroded zones in the side elevation and cross section at mid-height for all
three meshes are qualitatively similar to those for the CSCM model; see Figure 6-37b, Figure
6-38c and Figure 6-39c. The values of erosion strain computed using the CSCM model appear to
be transportable across concrete material models.
346

(a) 202020 mm mesh

(b) 404040 mm mesh

Figure 6-41 Maximum principal stress histories for 30 elements near or on the rear face of
the column, near its mid-height, CONCRETE_DAMAGE_REL3

347

(a) 202020 mm mesh

(b) 303030 mm mesh

(c) 404040 mm mesh


Figure 6-42 Internal energy histories for the concrete material models of the CSCM using
calculated erosion strains and an erosion algorithm of damage function, and
the CONCRETE_DAMAGE_REL3 model using calculated erosion strains

348

Side elevation

Cross section at mid-height


(a) 202020 mm mesh

Side elevation

Cross section at mid-height


(b) 303030 mm mesh

Side elevation

Cross section at mid-height


(c) 404040 mm mesh

Figure 6-43 Simulation resutls for the CONCRETE_DAMAGE_REL3 model

349

6.6.6.3 Concrete Compressive Strength


The influence of concrete compressive strength on damage, measured here using volume of
eroded material, is characterized by repeating some of the analysis described in prior sections
using 50 MPa concrete. The average peak tensile stress (traction) and traction rate for elements
near the rear face and mid-height of the column are approximately 20 MPa and 200 MPa/msec,
respectively. Per Figure 6-16, the corresponding erosion strains for the 20 20 20 mm,
303030 mm and 404040 mm meshes are 0.14, 0.010 and 0.008, respectively. Figure 6-44
shows internal energy histories for the 50 MPa concrete. For all three meshes, the internal energy
histories for the calculated erosion strains are similar to those for the damage-based erosion
algorithm. The damage, as described by eroded material and calculated using the damage-based
erosion algorithm (see Figures 6-45, 6-46 and 6-47), discrete values of erosion strain are similar
for the 30 and 40 mm meshes. The degree of damage is not substantially reduced by the use of
higher strength concrete; although the influence of concrete strength is likely masked somewhat
by the presence of transverse reinforcement.

350

(a) 202020 mm mesh

(b) 303030 mm mesh

(c) 404040 mm mesh


Figure 6-44 Internal energy histories for f c = 50 MPa

351

Side elevation

Cross section at mid-height


(a) Erosion strain=0.014

Side elevation

Cross section at mid-height


(b) Damage function

Figure 6-45 Simulation results for the 20 mm mesh at time = 25 msec, f c = 50 MPa

Side elevation

Cross section at mid-height


(a) Erosion strain=0.01

Side elevation

Cross section at mid-height


(b) Damage function

Figure 6-46 Simulation results for the 30 mm mesh at time = 25 msec, f c = 50 MPa
352

Side elevation

Cross section at mid-height


(a) Erosion strain=0.008

Side elevation

Cross section at mid-height


(b) Damage function

Figure 6-47 Simulation results for the 40 mm mesh at time = 25 msec, f c = 50 MPa
6.6.6.4 Confinement
Confinement of concrete is realized by transverse reinforcement, which is also used in reinforced
concrete columns to provide shear resistance. The vertical spacing of shear reinforcement in a
column often ranges between 25% and 50% of its effective depth. The spacing of the transverse
reinforcement for seismic applications may be as small as 100 mm. Closely spaced transverse
reinforcement is often specified for blast-resistant columns to better basket the core concrete and
increase the columns global and local deformation capacity.
Analyses are performed on a single cubical element and a model of a concrete cylinder to
demonstrate that the CSCM model for plain concrete adequately captures the influence of
passive confinement. An unconfined uniaxial compressive strength of 35.5 MPa is assumed. The
single element has dimensions of 202020 mm. Figure 6-48 shows the loading and boundary
conditions; the base of the element is free to expand. Lateral confining pressures, p, of 4, 8 and
16 MPa are applied to all four sides of the single element. The top of the single element is
subjected to a compressive, displacement-controlled loading, as shown in Figure 6-49. The

353

bottom of the element is uniformly supported. Figure 6-50 shows that the peak compressive
stress and ultimate strain increase with increasing confining pressure.
A standard 6-inch (152.4 mm) diameter by 12-inch (304.8 mm) tall concrete cylinder is analyzed
using the CSCM model and results are compared with those generated by the widely used
Mander model (Mander et al. 1988). Figure 6-51 shows the model used for analysis, which
invoked symmetry, allowing one quarter of the cylinder to be mode1ed; ux, uy and uz denote
translational displacement, and rx, ry and rz denote rotational displacements, in the x, y and z
directions, respectively. Uniform support is provided at the base of the model, with the base of
the model free to expand. An unconfined uniaxial compressive strength of 35.5 MPa is assumed.
Passive confining pressures of 4 and 8 MPa are imposed on the perimeter of the cylinder; results
are presented m Figure 6-52. Results of analysis using the Mander model are also presented in
the figures, where the uniaxial confined stress-strain relationships are calculated by substituting 4
and 8 MPa for the effective confining pressure, f l . Although there are differences between the
CSCM and Mander models, the results are quantitatively similar, and enable the use of the
CSCM model for considerations of the effect of passive confinement provided by transverse
reinforcement.
To understand the effect of transverse reinforcement (passive confinement) on damage, the
spacing of the transverse reinforcement assumed in prior sections (= 100 mm) is doubled to 200
mm or approximately 35% of the effective depth of the column. Simulation results for the
202020 mm, 303030 mm and 404040 mm meshes are presented in Figure 6-53, 6-54
and 6-55, respectively. As expected, the damage (or volume of eroded elements) increases with
the significant increase in the spacing of transverse reinforcement, as seen by comparing the
results in Figure 6-53 with those in panels (b) and (c) of Figure 6-37, the results in Figure 6-54
with those in panels (c) and (d) of Figure 6-38, and the results in Figure 6-55 with those in panels
(c) and (d) of Figure 6-39.

354

(a) Plan

(b) Elevation

Figure 6-48 Loading and boundary conditions of the single element for studying the effects
of confinement

Figure 6-49 Displacement-controlled loading at the top of the single element

355

Figure 6-50 Confinement effects on stress-strain relationships, 202020 mm element

ux = 0
ry = 0 304.8
rz = 0

ux = 0
ry = 0
rz = 0
76.2

76.2
uy = 0
rx = 0
rz = 0

76.2
uz = 0

(a) Plan

(b) Elevation

Figure 6-51 3D symmetry model of concrete cylinder (units: mm)

356

(a) Confining pressure, f l = 4 MPa

(b) Confining pressure, f l = 8 MPa


Figure 6-52 Compressive stress versus strain relationships

357

Side elevation

Cross section at mid-height


(a) Erosion strain=0.010

Side elevation

Cross section at mid-height


(b) Damage function

Figure 6-53 Simulation results for the 20 mm mesh with the spacing of transverse
reinforcement of 20 cm at time = 25 msec, f c = 35.5 MPa

Side elevation

Cross section at mid-height


(a) Erosion strain=0.008

Side elevation

Cross section at mid-height


(b)Damage function

Figure 6-54 Simulation results for the 30 mm mesh with the spacing of transverse
reinforcement of 20 cm at time = 25 msec, f c = 35.5 MPa
358

Side elevation

Cross section at mid-height


(a) Erosion strain=0.006

Side elevation

Cross section at mid-height


(b) Damage function

Figure 6-55 Simulation results for the 40 mm mesh with the spacing of transverse
reinforcement of 20 cm at time = 25 msec, f c = 35.5 MPa
6.6.6.5 Axial Load
The effect of axial loading due to gravity-load effects on the predicted damage to the column is
examined by applying axial static pressures of 0.1, 0.2 and 0.4 f c to the top of the column prior
to imposing the air-blast loading. A concrete strength of 35.5 MPa is assumed. The transverse
reinforcement is spaced at 100 mm over the height of the column. Two erosion strategies are
employed: calculated erosion strain and damage-based erosion algorithm. Results are presented
in Figures 6-56, 6-57 and 6-58 for the different mesh sizes; the case of zero axial load is included
for completeness. The column collapsed for all axial loadings and all mesh sizes due to extreme
damage in the cross section, principally at the mid-height of the column. The time to collapse
decreases with increasing axial load. It is important to note that if erosion is ignored, failure of
the column is not predicted, even for the axial pressure of 0.4 f c , which emphasizes the
importance of considering erosion when predicting damage to reinforced concrete columns.

359

Time = 25 msec

60 msec

90 msec

(a) Zero axial pressure, erosion strain=0.01

(b) Zero axial pressure, damage function

(c) Axial pressure: 0.1 f c , erosion strain=0.01

(d) Axial pressure: 0.1 f c , damage function


Figure 6-56 Simulation results for the 20 mm mesh with axial pressure loading, f c = 35.5
MPa

360

Time = 25 msec

60 msec

90 msec

(e) Axial pressure: 0.2 f c , erosion strain=0.01

(f) Axial pressure: 0.2 f c , damage function

(g) Axial pressure: 0.4 f c , erosion strain=0.01

(h) Axial pressure: 0.4 f c , damage function


Figure 6-56 Simulation results for the 20 mm mesh with axial pressure loading, f c = 35.5
MPa (cont.)

361

Time = 25 msec

60 msec

90 msec

160 msec

500 msec

(a) Zero axial pressure, erosion strain=0.008

(b) Zero axial pressure, damage function

(c) Axial pressure: 0.1 f c , erosion strain=0.008

(d) Axial pressure: 0.1 f c , damage function


Figure 6-57 Simulation results for the 30 mm mesh with axial pressure loading, f c = 35.5
MPa
362

Time = 25 msec

60 msec

90 msec

160 msec

500 msec

(e) Axial pressure: 0.2 f c , erosion strain=0.008

(f) Axial pressure: 0.2 f c , damage function

(g) Axial pressure: 0.4 f c , erosion strain=0.008

(h) Axial pressure: 0.4 f c , damage function


Figure 6-57 Simulation results for the 30 mm mesh with axial pressure loading, f c = 35.5
MPa (cont.)
363

Time = 25 msec

60 msec

90 msec

160 msec

500 msec

(a) Zero axial pressure, erosion strain=0.006

(b) Zero axial pressure, damage function

(c) Axial pressure: 0.1 f c , erosion strain=0.006

(d) Axial pressure: 0.1 f c , damage function


Figure 6-58 Simulation results for the 40 mm mesh with axial pressure loading, f c = 35.5
MPa

364

Time = 25 msec

60 msec

90 msec

160 msec

500 msec

(e) Axial pressure: 0.2 f c , erosion strain=0.006

(f) Axial pressure: 0.2 f c , damage function

(g) Axial pressure: 0.4 f c , erosion strain=0.006

(h) Axial pressure: 0.4 f c , damage function


Figure 6-58 Simulation results for the 40 mm mesh with axial pressure loading, f c = 35.5
MPa (cont.)

365

6.7 Values of Erosion Strain for Near-Field Blast Analysis


Step-by-step instructions are provided below to implement a unique value of erosion strain for
the near-field blast analysis of reinforced concrete components. The instructions are based on the
results of analysis presented in prior sections.
1. Prepare a numerical model of the component that explicitly models the vertical and transverse
reinforcement and the concrete. An element size of 40 mm is reasonable for the concrete.
2. Assign an appropriate constitutive model to the concrete (e.g., CSCM (Murray 2007),
CONCRETE_DAMAGE_REL3 (Malvar and Simons 1996) and CDP (Lee and Fenves, 1998))
and embed a limited number of linear elastic elements on and near the back face of the
component (expected to experience tensile stress upon wave reflection) to compute the
traction loading rate and maximum applied tensile stress.
3. Perform a simulation ignoring erosion to compute representative values of the traction loading
rate and maximum applied tensile stress on the elastic elements of step 2.
4. Using the representative values from step 3, select a value of erosion strain based on concrete
strength (proximity to either 35 MPa or 50 MPa). Set the peak principal tensile strain to the
erosion strain and perform the blast simulation.

366

SECTION 7
SUMMARY AND CONCLUSIONS
7.1 Summary
Air-blast loadings for security design of geometrically simple, components and structures are
typically calculated using empirical charts available in textbooks and US government-published
documents such as TM 5-1300 (Department of the Army, Navy and Air Force 1969 and 1990),
TM 5-858-3 (DoA 1984), UFC 3-340-02 (DoD 2008), Smith and Hetherington (1994), Cormie et
al. (2009), Dusenberry (2010) and ASCE (2011). These charts are based on studies performed by
Kingery and his co-workers (e.g., Kingery and Bulmash, 1984). Computational fluid dynamics
(CFD) codes such as CTH (McGlaun et al. 1990), Air3d (Rose 2006), AUTODYN (ANSYS
2013a) and LS-DYNA (LSTC 2013) can also be used to calculate air-blast loadings on
geometrically simple shapes but are preferred for complex shapes and geometries such as urban
environments and streetscapes.
The empirical charts of Kingery and Bulmash, and their underlying polynomials, were
constructed around measured and inferred data. These charts have not been validated in the near
field because pressures and temperatures are very high. Although commercially available
pressure transducers are now capable of measuring pressures expected in the near field, they
cannot sustain the co-existing temperatures of thousands Kelvin. The only practical strategy to
confirm pressures and impulses in the near field, both incident and reflected, is to use verified
and validated CFD codes.
The goal of this report was to characterize the effects of detonations of ideal high explosives and
their influence on geometrically simple structures and components, for the purpose of informing
blast-resistant design. The near-field region was emphasized because there are virtually no
measured data available for validation. This report has provided 1) validated methodologies and
guidance for modeling detonations of high explosives, 2) an assessment of the accuracy of the
widely used empirical charts for air-blast loading through CFD analysis, 3) new polynomials and
design charts for air-blast loadings suitable for inclusion in design standards, 4) an assessment of
367

the reflection coefficients presented in government manuals and textbooks for peak overpressure,
and 5) a detailed discussion of the blast analysis of a sample reinforced concrete column and the
development of a technical basis for selecting values of erosion strain.
AUTODYN was verified and validated in Chapter 2 for 1D and 2D air-blast calculations through
1) 1D simulations and comparisons of results with those presented by Needham (2010), who
performed numerical and first-principles calculations for near- and intermediate-field detonations,
2) cross-code comparisons in 1D and 2D using CTH and LS-DYNA for far-field detonations, 3)
2D simulations and comparisons of results to measurements reported by Goodman (1960) and
Huffington and Ewing (1985) for near-field detonations, and 4) 2D simulations and comparisons
of results to measurements reported by Frost et al. (2008) for far-field detonations. Incident and
reflected overpressure histories were simulated to examine the influences of mesh size,
expansion of detonation products and afterburning, temperatures near the charge face, and the
efficacy of reflecting and transmitting boundaries.
Chapter 3 presented the results of AUTODYN studies of air-blast parameters over a wide range
of scaled distance. Incident and normally reflected peak overpressures and impulses, and shock
front arrival times, were compared with predictions from the Kingery and Bulmash (or UFC 3340-02) charts.
Chapter 4 proposed new polynomials and design charts, based on the CFD calculations of
Chapter 3, for air-blast parameters over a wide range of scaled distance, namely, 0.0553 Z 40
m/kg1/3 (0.139 Z 100 ft/lb1/3).
Overpressure reflection coefficients and reflected scaled impulses as a function of the angle of
incidence for airblast calculations were studied in Chapter 5, using a verified CFD code,
AUTODYN, and using data reported in TM 5-858-3 (DoA 1984) and UFC 3-340-02 (DoD
2008). The study was motivated by the significant differences in the coefficients in the Mach
reflection region between TM 5-858-03 and UFC 3-340-02. Incident and reflected peak
overpressures and impulses were simulated using AUTODYN for 0.16 Z 8 m/kg1/3 and
angles of incidence between zero and ninety degrees. The effects of Mach reflection and

368

expanding detonation product on the reflection coefficient and reflected scaled impulse were
examined.
Chapter 6 presented analytical studies performed to characterize erosion strain in concrete for the
purpose of damage analysis. Single element simulations were undertaken to characterize erosion
strain as a function of variables such as concrete strength, rate of traction loading, maximum
traction loading, and size of finite element. Finite element analysis of a sample reinforced
concrete column was performed using alternate representations of erosion strain and the damagebased erosion algorithm developed by Murray (2007). Other factors that affect the response of a
reinforced concrete column, including concrete compressive strength, volume of transverse
reinforcement, and co-existing axial (gravity) load, were studied.

7.2 Conclusions
The key findings and recommendations in this report are:
1. AUTODYN produces similar predictions of 1) hydrodynamic parameters to those of
Needham (2010) for near-field detonations, 2) incident overpressure and impulse histories to
those obtained using CTH and LS-DYNA for far-field detonations, 3) reflected scaled
impulses to measurements reported by Goodman (1960) and Huffington and Ewing (1985)
for near-field detonations, and 4) normally reflected peak overpressures and impulses to
measurements of Frost et al. (2008) for far-field detonations. AUTODYN is verified and
validated for near- and far-field air-blast calculations in 1D and 2D.
2. For CFD calculations of incident and reflected overpressure histories, a trial cell size of
R/500 is sufficient for Z 8 m/kg1/3, where R is the distance from the center of the charge to
the monitoring location and Z is scaled distance. The choice of cell size must also allow the
accurate meshing of the charge.
3. The widely used Friedlander equation does not capture the shape of the overpressure history
in the near field due to the expansion of detonation products, afterburning, and Mach stem
formation. CFD tools must be used to characterize overpressure histories in the near field.

369

The Friedlander equation can be used to characterize incident and normally reflected
overpressure histories for Z 0.83 m/kg1/3.
4. Afterburning has no effect on peak overpressure but can increase impulse if sufficient
oxygen is present, the available oxygen mixes well with the fuels, and the temperature is
greater than 1800 K (for TNT). The effect of afterburning is likely small in the near field.
5. Transmitting boundaries may partially reflect incident pressures in 3D models. The location
of a perpendicular transmitting boundary in a 3D problem should be established using a
process similar to that reported in Section 2.6. A minimum perpendicular dimension equal to
the charge radius is recommended.
6. AUTODYN simulations predict values of incident peak overpressure, incident impulse, and
normally reflected peak overpressure that are much greater than those presented in UFC 3340-02 for Z < 0.08, 0.1, and 0.4 m/kg1/3, respectively. UFC 3-340-02 overpredicts the
AUTODYN calculations of incident impulse for 0.1 < Z < 40 m/kg1/3 by 10% to 25%, which
is consistent with the observations of Bogosian et al. (2002).
7. The design polynomials and charts of Chapter 4 enable good predictions of incident and
normally reflected peak overpressure and impulse across a wide range of scaled distance,
namely, 0.0553 Z < 40 m/kg1/3.
8. For Z < 0.4 m/kg1/3, the AUTODYN-calculated reflection coefficients are much greater than
the values presented in UFC 3-340-02, although this may be of little practical importance
because CFD tools should be used in the near field to estimate the overpressure histories for
design.
9. For Z 0.4 m/kg1/3 (1 ft/lb1/3) (or Ps approximately 5.6 MPa [815 psi]) and 40o, the
overpressure reflection coefficients calculated using AUTODYN are similar to those reported
in both UFC 3-340-02 and TM 5-858-3. For Z 0.4 m/kg1/3 and > 40o, AUTODYN
calculates similar values of overpressure reflection coefficients to those in UFC 3-340-02; the
corresponding chart in TM 5-858-3 does not capture the effect of Mach reflection.

370

10. The use of an erosion strain that is independent of concrete strength, rate of traction loading,
maximum tensile traction and element size is inappropriate for damage and/or failure
calculations. For 35.5 MPa and 50 MPa concrete, and a 40 mm element size or smaller, the
erosion strain varies between 0.005 and 0.015 for maximum applied tensile tractions ranging
between 10 MPa and 40 MPa and loading traction rates of between 50 MPa/msec and 250
MPa/msec. The use of an unreasonably high value of erosion strain will underestimate the
volume of eroded material and underpredict the likelihood of gravity-load collapse.
11. Given that the maximum aggregate size in a reinforced concrete column will typically exceed
20 mm, an element size of 40 mm appears reasonable for the finite element analysis of a
reinforced concrete column.
12. The use of the damage-based erosion algorithm in LS-DYNA resulted in similar predictions
of damage, including volume of eroded material, to that predicted by analysis using the stepby-step procedure and a robust value for the erosion strain. There is no benefit to using
erosion strain for analysis if a robust damage-based erosion algorithm is available.

371

SECTION 8
REFERENCES
ACI. 2011. Building Code Requirements for Structural Concrete (ACI 318-11) and Commentary.
American Concrete Institute, Farmington Hills, MI.
ASCE. 2011. Blast Protection of Buildings (ASCE/Structural Engineering Istitute (SEI) 59-11).
American Society of Civil Engineers, Reston, VA.
ASME. 2006. Guide for Verification and Validation in Computational Solid Mechanics. ASME
V & V 10-2006, American Society of Mechanical Engineers, New York, NY.
ASTM. 2003. Standard Specification for Deformed and Plain Billet-Steel Bars for Concrete
Reinforcement (ASTM A 615/A 615M-03a). American Society of Testing and Materials,
Philadelphia, PA.
Anderson, J. D. 2003. Modern Compressible Flow: with Historical Perspective, 3rd edition.
McGraw-Hill, Boston, MA.
ANSYS. 2005. Remapping Tutorial, Revision 4.3. ANSYS, Inc., Canonsburg, PA.
ANSYS. 2013a. AUTODYN Users Manual Version 15.0. ANSYS, Inc., Canonsburg, PA.
ANSYS. 2013b. Personal Communication. ANSYS, Inc., Canonsburg, PA.
Baker, W. E. 1973. Explosions in Air. University of Texas Press, Austin, TX.
Baker, W. E., Cox, P. A., Westine, P. S., Kulesz, J. J. and Strehlow, R. A. 1983. Explosion
Hazards and Evaluation. Elsevier Scientific Publishing Company, Amsterdam; Oxford; New
York.
Ballantyne, G. J., Whittaker, A. S., Aref, A. J. and Dargush, G. F. 2009. Air-blast effects on
structural shapes. Technical Report MCEER-09-0002, The State University of New York at
Buffalo, Buffalo, NY.
373

Baant, Z. P. 1986. Mechanics of distributed cracking. Journal of Applied Mechanics 39 (5):


675-705.
Baant, Z. P. and Chang, T. P. 1987. Nonlocal finite-element analysis of strain-softening solids.
Journal of Engineering Mechanics 113 (1): 89-105.
Baant, Z. P. and Pijaudier-Cabot, G. 1988. Nonlocal continuum damage, localization instability
and convergence. Journal of Applied Mechanics 55 (2): 287-293.
Belytschko, T., Baant, Z. P., Hyun, Y. W. and Chang, T. P. 1986. Strain-softening materials and
finite-element solutions. Computers & Structures 23 (2): 163-180.
Biggs, J. M. 1964. Introduction to Structural Dynamics. McGraw-Hill, NY.
Bogosian, D., Ferritto, J., and Shi, Y. 2002. Measuring uncertainty and conservatism in
simplified blast models. Proceedings, 30th Explosives Safety Seminar, Atlanta, Georgia.
Boris, J. P and Book, D. L. 1971. Flux-correct-transport: I. SHASTA, a fluid transport algorithm
that works. Journal of Computational Physics 11: 38-69.
Borve, S., Bjerke, A., Omang, M. and Truslen, J. 2008. A comparative study of ANSYS
AUTODYN and RSPH simulations of blast waves. Proceedings, 3rd ERCOFTAC SPHERIC
Workshop on SPH applications, Lausanne, Switzerland.
Brvik, T., Hopperstad, O., Berstad, T. and Langseth, M. 2001. A computational model of
viscoplasticity and ductile damage for impact and penetration. European Journal of MechanicsA/Solids 20 (5): 685-712.
Brannon, R. M. and Leelavanichkul, S. 2009. Survey of four damage models for concrete.
Sandia Report No. SAND2009-5544. Sandia National Laboratories, University of Utah,
Albuquerque, NM ; Livermore, CA.
Brara, A. and Klepaczko, J. R. 2007. Fracture energy of concrete at high loading rates in tension.
International Journal of Impact Engineering 34 (3): 424-435.

374

Brode, H. L. 1955. Numerical solution of spherical blast waves. Journal of Applied Physics 26
(6): 766-776.
Browning, R. S., Sherburn, J. A. and Schwer, L. E. 2013. Predicting blast loads using LS-DYNA
and CTH. Proceedings, ASCE Structures Congress, Pittsburgh, PA.
Chen, W.-F. 2007. Plasticity in Reinforced Concrete. J. Ross Publishing Inc., Fort Lauderdale, Fl.
Chopra, A. K. 2012. Dynamics of Structures: Theory and Applications to Earthquake
Engineering, 4th edition. Prentice Hall, Upper Saddle River, NJ.
Comit euro-international du bton (CEB). 1993. CEB-FIP Model Code 1990: Design Code. T.
Telford, London.
Cook, R. D. 2002. Concepts and Applications of Finite Element Analysis, 4th edition. Wiley,
New York, NY.
Cormie, D., Mays, G. and Smith, P. 2009. Blast Effects on Buildings, 2nd edition. London:
Thomas Telford.
Cormie, D., Wilkinson, W., Shin, J. and Whittaker, A. S. 2013. Scaled-distance relationships for
close-in detonations. Proceedings, 15th International Symposium on Interaction of the Effects of
Munitions with Structures, Potsdam, Germany.
Coughlin, A. M., Musselman, E. S., Schokker, A. J. and Linzell, D. G. 2010. Behavior of
portable fiber reinforced concrete vehicle barriers subject to blasts from contact charges.
International Journal of Impact Engineering 37 (5): 521-529.
Courant, R. and Friedrichs, K. O. 1948. Supersonic Flow and Shock Waves. Interscience
Publishers, New York, NY.
Cullis, I. G. and Huntington-Thresher, W. 2007. Blast structure interaction and the role of
secondary combustion. Proceedings, 23rd International Symposium on Ballistics, Tarragona,
Spain.

375

Danielson, K. T., ODaniel, J. L., Akers, S. A. and Adley, M. D. 2008. Numerical aspects and
procedures for modeling reinforced concrete structures under extreme impulsive loadings.
Materials under Extreme Loadings - Application to Penetration and Impact. Proceedings, 2nd
US-France Conference Organized by the International Center for Applied Computational
Mechanics. Rocamadour, France.
Davison, L. 2008. Fundamentals of Shock Wave Propagation in Solids. Springer, NY.
DoA. 1984. Designing Facilities to Resist Nuclear Weapons Effects: Structures (TM 5-858-3).
Department of the Army, Headquarters, United States. Washington, DC.
DoA. 1986. Fundamentals of Protective Design for Conventional Weapons (TM 5-855-1).
Department of the Army, Headquarters, United States. Washington, DC.
Department of the Army, Navy and Air Force. 1969. Structures to Resist the Effects of
Accidental Explosions (with Addenda). Army Technical Manual (TM 5-1300), Navy Publication
(NAVFAC P-397), Air Force Manual (AFM 88-22). Washington, DC.
Department of the Army, Navy and Air Force. 1990. Structures to Resist the Effects of
Accidental Explosions (with Addenda). Army Technical Manual (TM 5-1300), Navy Publication
(NAVFAC P-397), Air Force Manual (AFM 88-22), Revision 1. Washington, DC.
DoD. 2008. Unified Facilities Criteria (UFC): Structures to Resist the Effects of Accidental
Explosions (UFC 3-340-02). Departments of Defense, Washington, DC.
Doan, L. R. and Nickel, G. H. 1963. A subroutine for the equation of state of air. RTD (WLR)
TN63-2. Air Force Weapons Laboratory.
Dobratz, B. M. and Crawford, P. C. 1985. LLNL Explosive Handbook, Properties of chemical
explosives and explosive simulants. Report UCRL-52997, Lawrence Livermore National
Laboratory, Livermore, CA.
Donahue, L. K. 2008. Afterburning of TNT detonation in air. M.A.Sc. Thesis. Dalhousie
University, Halifax, Nova Scotica.
376

Dusenberry, D. O. 2010. ed. Handbook for Blast-Resistant Design of Buildings. Wiley, Hoboken,
NJ.
Fawaz, Z., Zheng, W. and Behdinan, K. 2004. Numerical simulation of normal and oblique
ballistic impact on ceramic composite armours. Composite Structures 63 (3-4): 387395.
Federal Emergency Management Agency (FEMA). 2003. Reference Manual to Mitigate
Potential Terrorist Attacks against Buildings. FEMA-426, Washing DC.
Fickett, W. and Davis W. C. 1979. Detonation: Theory and Experiment. Dover Publications Inc.,
Mineola, NY.
Frost, D.L., Cairns, M., Goroshin, S., Leadbetter, J., Ripley, R., and Zhang, F. 2008. Reflected
heterogeneous blast. Proceedings, 20th International Symposium on Military Aspects of Blast
and Shock, Oslo, Norway.
Godunov, S. K. 1959. A difference scheme for numerical solution of discontinuous solution of
hydrodynamic equations. Math. Sbornik, 47, 271-306, translated US Joint Publ. Res. Service,
JPRS 7226, 1969.
Goodman, H. J. 1960. Complied free-air blast data on bare spherical Pentolite. BRL Report No.
1092, US Army Ballistic Research Laboratory, Aberdeen Proving Ground, MD.
Gray, G. T., Chen, S. R., Wright, W. and Lopez, M. F. 1994. Constitutive equations for annealed
metals under compression at high strain rates and high temperatures. LA-12669-MS. Los Alamos
National Laboratory, Los Alamos, NM.
Griffith, W. C. and Bleakney, W. 1954. Shock waves in gases. American Journal of Physics 22
(9): 597-612.
Grote, D. L., Park, S. W. and Zhou, M. 2001. Dynamic behavior of concrete at high strain rates
and pressures: I. experimental characterization. International Journal of Impact Engineering 25
(9): 869-886.

377

Hao, H. and Zhou, X. Q. 2007. Concrete material model for high rate dynamic analysis.
Proceedings, 7th International Conference on Shock and Impact Loads on Structures. Beijing,
China. p. 753-768.
Henrych, J. 1979. The Dynamics of Explosion and Its Use. Elsevier Scientific Pub. Co.,
Amsterdam ; New York.
Huffington, Jr., N. J. and Ewing, W. O. 1985. Reflected impulse near spherical charges.
Technical Report BRL-TR-2678, US Army Ballistic Research Laboratory, Aberdeen Proving
Ground, MD.
Hyde, D. W. 1992. ConWep: Conventional Weapons Effects (Application of TM 5-855-1). US
Army Corps of Enigneers, Waterways Experiment Station, Vicksburg, MS.
Islam, M. J., Liu, Z. and Swaddiwudhipong, S. 2011. Numerical study on concrete
penetration/perforation under high velocity impact by ogive-nose steel projectile. Computers and
Concrete 8 (1): 111-123.
Ivanov, M. S., Vandromme, D., Fomin, V. M., Kudryavtsev, A. N., Hadjadj, A., and
Khotyanovsky, D. V. 2011. Transition between regular and Mach reflection of shock waves: new
numerical and experimental results. Shock Waves 11: 199207
Johnson, G. R. and Cook, W. H. 1983. A constitutive model and data for metals subjected to
large strains, high strain rates and high temperatures. Proceedings, 7th International Symposium
on Ballistics. International Ballistics Committee, The Hague, Netherlands. p. 541-547.
Johnson, O. T., Patterson, II, J. D. and Olson, W. C. 1957. A simple mechanical method for
measuring the reflected impulse of air blast waves. BRL Memorandum Report No. 1088, US
Army Ballistic Research Laboratory, Aberdeen Proving Ground, MD.
Kingery, C. N. 1966. Air blast parameters versus distance for hemispherical TNT surface bursts.
Report No. 1344, Ballistic Research Laboratories, Aberdeen Proving Ground, MD.

378

Kingery, C. N. and Bulmash, G. 1984. Airblast parameters from TNT spherical air burst and
hemispherical surface burst. Report ARBRL-TR-02555, US Army Ballastic Research Laboratory,
Aberdeen Proving Ground, MD.
Kingery, C. N. and Pannill, B. F. 1964. Peak overpressure vs scaled distance for TNT surface
bursts (hemispherical charges). BRL Memorandum Report 1518, Ballistic Research Laboratories,
Aberdeen Proving Ground, MD.
Kinney, G. F. 1962. Explosive Shocks in Air. The Macmillan Company, New York, NY.
Kinney, G. F. and Graham, K. J. 1985. Explosive Shocks in Air. 2nd edition. Springer-Verlag,
New York, NY.
Kolsky, H. 1953. Stress Waves in Solids. Clarendon Press Oxford.
Krauthammer, T. 2008. Modern Protective Structures. CRC Press, Boca Raton, FL.
Kurtaran, H., Buyuk, M. and Eskandarian, A. 2003. Ballistic impact simulation of GT model
vehicle door using nite element method. Theoretical and Applied Fracture Mechanics 40 (2):
113-121.
Lambert, D. E. and Ross, C. A. 2000. Strain rate effects on dynamic fracture and strength.
International Journal of Impact Engineering 24 (10): 985-998.
Lee, J and Fenves, G. 1998. Plastic-damage model for cyclic loading of concrete structures.
Journal of Engineering Mechanics, 124 (8): 892-900.
Lignos, D. G., Chung, Y., Nagae, T. and Nakashima, M. 2011. Numerical and experimental
evaluation of seismic capacity of high-rise steel buildings subjected to long duration earthquakes.
Computers & Structures 89 (11-12): 959-967.
LSTC. 2013. LS-DYNA Keyword User's Manual Ver. R7.0. Livermore Software Technology
Corporation, Livermore, CA.

379

Lu, Y. and Xu, K. 2004. Modelling of dynamic behaviour of concrete materials under blast
loading. International Journal of Solids and Structures 41 (1): 131-143.
Luccioni, B. M., Ambrosini, R. D. and Danesi, R. F. 2004. Analysis of building collapse under
blast loads. Engineering Structures 26 (1): 63-71.
Malvar, L. J. and Ross, C. A. 1998. Review of strain rate effects for concrete in tension. ACI
Materials Journal 95 (6): 735-739.
Malvar, L. J. and Simons, D. 1996. Concrete material modeling in explicit computations.
Proceedings, Workshop on Recent Advances in Computational Structural Dynamics and High
Performance Computing, USAE Waterways Experiment Station, Vicksburg, MS, p. 165-194
Mander, J. B., Priestley, M. J. N. and Park, R. 1988. Theoretical stress-strain model of confined
concrete. Journal of Structural Engineering 114 (8): 1804-1826.
McGlaun, J. M., Thompson, S. L. and Elrick, M. G. 1990. CTH: a three-dimensional shock wave
physics code. International Journal of Impact Engineering 10 (1-4): 351-360.
McNesby, K. L., Homan, B. E., Ritter, J. J., Quine, Z., Ehlers, R. Z. and McAndrew, B. A. 2010.
Afterburn ignition delay and shock augmentation in fuel rich solid explosives. Propellants,
Explosives, Pyrotechnics 35 (1): 57-65.
Mehanny, S. S. and Deierlein, G. G. 2000. Modeling and assessment of seismic performance of
composite frames with reinforced concrete columns and steel beams. Report No. 135. John A.
Blume Earthquake Engineering Research Center, Stanford, CA.
Meyers, M. A., Murr, L. E. and Staudhammer, K. P. 1992. Shock-Wave and High-Strain-Rate
Phenomena in Materials. CRC Press, NY.
Mouton, C. A. 2007. Transition between regular and Mach reflection in the dual-solution domain.
PhD. Dissertation, California Institute of Technpology, Pasadena, CA.

380

Murray, Y. D. 2007a. Evaluation of LS-DYNA Concrete Material Model 159. Report No.
FHWA-HRT-05-063. Federal Highway Administration, DC.
Murray, Y. D. 2007b. Users Manual for LS-DYNA Concrete Material Model 159. Report No.
FHWA-HRT-05-063. Federal Highway Administration, DC.
Needham, C. E. 2010. Blast Waves. Springer, NY.
Ngo, T., Lumantarna, R., Whittaker, A. S. and Mendis, P. 2014. Arena testing using a large high
explosive, submitted to Journal of Structural Engineering.
Norris, C. H., Hansen, R. J., Holley, JR. M. J., Biggs, J. M., Namyet, S. and Minami, J. K. 1959.
Structural Design for Dynamic Loads. McGraw-Hill, New York; Toronto; London.
Oberkampf, W. L. and Trucano, T. G. 2002. Verification and validation in computational fluid
dynamics. SAND2002-0529, Sandia National Laboratories, Albuquerque, NM.
Polanco-Loria, M., Hopperstad, O. S., Brvik, T. and Berstad, T. 2008. Numerical predictions of
ballistic limits for concrete slabs using a modified version of the HJC concrete model.
International Journal of Impact Engineering 35 (5): 290-303.
Popovics, S. 1970. A review of stress-strain relationships for concrete. ACI Journal 67 (3): 243248.
Ragueneau, F. and Gatuingt, F. 2003. Inelastic behavior modelling of concrete in low and high
strain rate dynamics. Computers & Structures 81 (12): 1287-1299.
Ritzel, D. V. and Matthews, K. 1997. An adjustable explosion-source model for CFD blast
calculations. Proceedings, 21st International Symposium on Shock Waves, Great Keppel Island,
Australia, 1: 97-102.
Rose, T. A. 2006. A Computational Tool for Airblast Calculations - Air3D version 9 users' guide.
Engineering Systems Department, Cranfield University, United Kingdom.

381

Rubin, M. B. 1991. Simple, convenient isotropic failure surface. Journal of Engineering


Mechanics 117 (2): 348-369.
Ruiz, G., Zhang, X. X., Tarifa, M., Yu, R. C. and Camara, M. 2009. Fracture energy of highstrength concrete under different loading rates. Anales de Mecnica de la Fractura 26 2: 513518.
Sandler, I. S. and Wright, J. P. 1984. Summary of strain-softening. Proceedings, DARPA-NSF
Workshop on the Theoretical Foundations for Large-Scale Computations for Nonlinear Material
Behavior. Edited by Nemat-Nasser, S. Northwestern University, Evanston, IL. p. 285-315.
Schuler, H., Mayrhofer, C. and Thoma, K. 2006. Spall experiments for the measurement of the
tensile strength and fracture energy of concrete at high strain rates. International Journal of
Impact Engineering 32 (10): 1635-1650.
Schwer, D. 2008. Regular and Mach reflections to Mach 18 with air and TNT detonation
products. NRL/MR/641008-9132, Naval Research Laboratory, Washington, DC.
Schwer, L. E. 2004. Preliminary assessment of non-Lagrangian methods for penetration
simulation. Proceedings, 8th International LS-DYNA Users Conference. Dearborn, MI. p. 8.18.12.
Sherkar, P., Whittaker A. S. and Aref, A. J. 2010. Modeling the effects of detonations of high
explosives to inform blast-resistant design. Technical Report MCEER-10-0009, The State
University of New York at Buffalo, Buffalo, NY.
Smith, P. D. and Hetherington, J. G. 1994. Blast and Ballistic Loading of Structures.
Butterworth-Heinemann, Oxford; Boston.
Souers, P. C., Forbes, J. W., Fried, L. E., Howard, W. M., Anderson, S., Dawson, S., Vitello, P.
and Garza, R. 2001. Detonation energies from the cylinder test and CHEETAH V3.0.
Propellants, Explosives, Pyrotechnics 26 (4): 180-190.

382

Swisdak, M. M., Jr. 1994. Simplified Kingery airblast calculations. Naval Surface Warfare
Center, Indian Head Division, Silver Spring, MD.
Tan, P., Lee, B. and Tsangalis, C. 2009. FEA modelling prediction of the transmitted
overpressure and particle acceleration within a frame subjected to shock tube blast loadings.
Proceedings, 18th World IMACS / MODSIM Congress. Cairns, Australia. p. 1657-1663.
Teng, T. L., Chu, Y. A., Chang, F. A. and Chin, H. S. 2004. Simulation model of impact on
reinforced concrete. Cement and Concrete Research 34 (11): 2067-2077.
Timoshenko, S. P. and Goodier, J. N. 1970. Theory of Elasticity, 3rd edition. McGraw-Hill, NY.
Tuler, F. R. and Butcher, B. M. 1984. A criterion for the time-dependence of dynamic fracture.
International Journal of Fracture 26 (4): 322-328.
Wight, J. K. and MacGregor, J. G. 2011. Reinforced Concrete: Mechanics and Design, 6th
edition. Prentice Hall, Upper Saddle River, NJ.
Wu, C., Oehlers, D. J., Rebentrost, M., Leach, J., Whittaker, A. S. 2009. Blast testing of ultrahigh performance fibre and FRP-retrofitted concrete. Engineering Structures 31 (9): 2060-2069.
Xu, K. and Lu, Y. 2006. Numerical simulation study of spallation in reinforced concrete plates
subjected to blast loading. Computers & Structures 84 (5-6): 431-438.
Xu, S., Zhao, Y. and Wu, Z. 2006. Study on the average fracture energy for crack propagation in
concrete. Journal of Materials in Civil Engineering 18 (6): 817-824.
Zhou, X. Q., Kuznetsov, V. A., Hao, H. and Waschl, J. 2008. Numerical prediction of concrete
slab response to blast loading. International Journal of Impact Engineering 35 (10): 1186-1200.
Zukas, J. A. 1990. High Velocity Impact Dynamics. Wiley, New York, NY.

383

APPENDIX A
SPEED OF SOUND
A.1 Introduction
The speed of sound is a pressure disturbance that propagates through a medium. In a solid, stress
or pressure waves propagate at the speed of sound because particles oscillate around their
original positions due to strong intermolecular forces (attractions) that hold adjacent particles
together. In a fluid, materials move with time because intermolecular forces are weak. Pressure
waves in a fluid propagate at the sum of the speed of sound in the medium (e.g., air) and the
particle velocity.

A.2 Determination of Speed of Sound in Solids and Fluids


For an elastic solid with a constant bulk modulus, the speed of sound, csolid , in the longitudinal
direction is (Kolsky 1953)
csolid

(A-1)

where E is the Youngs modulus and is density. In a fluid, the speed of sound, c fluid , is given
by (Kinney and Graham 1985)
c fluid

dp
or
d

(A-2)

where p is pressure and K is the bulk modulus, which is volumetric elasticity and an extension of
the Youngs modulus, E, to three dimensions. The bulk modulus is calculated as
K v

dp
dp
or
dv
d

(A-3)

For an ideal gas, the formula of the speed of sound in a fluid in Equation A-2 is further
simplified using entropy, which is defined as a measure of disorder or dissipated energy in a
system. For an idea gas, the entropy, s, remains constant, namely,

385

ds 0 cv

dT
1
dT 1
pd ( ) c p
dp
T

(A-4)

where cv is the specific heat at constant volume, c p is the specific heat at constant pressure, T is
temperature. Equation A-4 can be rewritten for dT/T as a function of other variables:

cp
dT c p pd cv dp
dp p

or
(
)

2
T

cv
d

(A-5)

Using Equation A-5, the speed of sound in a fluid for an ideal gas can be simplified:
c fluid

dp
p

RT ( p RT )
d

(A-6)

where R is the gas constant.

386

APPENDIX B
DETONATION WAVE AND PRESSURE
B.1 Introduction
The pressure after detonation in a high explosive is well described by the Chapman-Jouguet (CJ)
conditions (e.g., Fickett and David 1979, Smith and Hetherington 1994), which are a restatement
of the Rankine-Hugoniot relationships for one-dimensional (1D) steady11 and inviscid12 flow. A
detonation (shock) wave propagates at constant velocity through an unreacted high explosive.
The velocity at the detonation front is equal to the sum of the speed of sound and the local
particle velocity immediately behind the detonation front. A thin chemical reaction zone follows
the front of the detonation wave, which is explained by the Zeldovich-von Neumann-Dring
(ZND) theory (e.g., Davison 2008). The detonation wave compresses the charge to a high
pressure termed the von Neumann spike, at which the charge is still unreacted. The spike
indicates the onset of chemical reactions. The chemical reactions terminate with chemical
equilibrium 13 immediately behind the reaction zone, which is defined as the CJ plane. The
pressure measured at the CJ plane is termed the CJ pressure or the detonation pressure. Figure B1 shows pressures and other states across the shock front. The ZND theory assumes that the
reaction zone is steady; thermodynamic states (e.g., pressure, density, internal energy) in the
reaction zone are maintained constant by the chemical reactions. The region behind the CJ plane
is unsteady, where the states vary continuously with time. The unsteady flow is described by a
Taylor wave (e.g., Davison 2008), which is a rarefaction wave centered at the origin of the
explosive. The centered rarefaction wave expands at the velocity equal to the sum of the sound
speed and the local flow (particle) velocity and the wave expansion decreases pressures and

11

In steady flow, all conditions at any point are constant with respect to time, whereas in unsteady flow,
all conditions at any point are not constant with respect to time.
12
Inviscid flow assumes no viscosity in the flow. The viscous effects can be neglected for a gas in a
supersonic flow resulted from detonation.
13
Chemical equilibrium is the state in which the forward and backward reactions proceed at the same rate,
or reactions halt. There are thus no net changes in the concentrations of reactants and products (Anderson
2003).

387

temperatures in the region behind the CJ plane. When the detonation wave exits the surface of
the charge, another rarefaction wave is created and propagates toward the origin of the charge at
the speed of sound minus the local particle velocity. The relationships of the CJ detonation
pressure and states across the detonation front are formulated using the Rankine-Hugoniot
equations in the following section.

Figure B-1 States across a detonation (shock) wave in unsteady flow (adapted from Fickett
and Davis 1979)
B.2 Hugoniot Curves and Rayleigh Line for the Detonation Front
The Rankine-Hugoniot equations in a steady and inviscid flow, conserving mass, momentum and
energy, are

1u1 0u0

(B-1)

p1 1u12 p0 0u02

(B-2)

1
1
h1 u12 h0 u02
2
2

(B-3)

where is density, u is particle velocity, p is pressure, h is enthalpy, and subscripts 1 and 0


denote the compressed and undisturbed media, respectively, immediately behind and ahead of
388

the detonation front, respectively. Figure B-1 shows the states across the detonation front in
unsteady flow, which is transformed to steady flow in Figure B-2. Equations are derived across
the detonation front using the Rankin-Hugoniot equations, CJ conditions, and steady-state
variables of Figure B-2. Equation B-1 can be rewritten by replacing u1 and u0 by D u p and

D , respectively,

1 ( D u p ) 0 D

(B-4)

where D is the detonation velocity, u p is particle velocity, in the unsteady flow, immediately
behind the detonation front. Similar to Equation B-4, Equation B-2 is expressed as

p1 p0 0 Du p

(B-5)

The enthalpy, h, is a measure of energy effects for systems and also called total energy. The
enthalpy is calculated as the sum of the internal energy and the pressure-volume product:

h pv e p / e

(B-6)

where v is the specific volume and e is the specific internal energy. Substituting Equation B-6
into Equation C-3 yields
p1

(u p D)2
2

e1

p0

D2
e0
2

(B-7)

Equations B-4 and B-5 can be rewritten for u p as

up

D( 1 0 )
1

(B-8)

up

p1 p0
0 D

(B-9)

From Equations B-5, B-7 and B-8,

u p2

( 1 0 )( p1 p0 )
10

Du p

(B-10)

p1 p0
0

(B-11)

389

Substituting Equations B-10 and B-11 into Equation B-7 to eliminate the parameters, u p and D,
yields

1
1 1
1
e1 e0 ( p1 p0 )( ) or ( p1 p0 )(v0 v1 )
2
0 1
2

(B-12)

The p-v diagram using this equation is called the Hugoniot curve for unreacted explosives. To
produce the Hugoniot curve for the reacted explosives, the energy of chemical reactions, Q,
should be added to the equation, such that e1 is replaced by e1 Q .
The equations of conservations of mass and momentum in Equations B-4 and B-5, respectively,
are reorganized to give
1
1
p1 p0 0 2 D 2 or 0 2 D 2 v0 v1
0 1

(B-13)

This is the equation of the Rayleigh line, which is a straight line with slope of 02 D 2 . The
Hugoniot curve and the Rayleigh line are illustrated in Figure B-3, where p p1 and v v1 . The
point ( p0 , v0 ) is the origin, which is the initial state of an unreacted explosive. The Rayleigh line
passes through the origin and always has a negative slope of 02 D 2 as shown in Equation B-13,
such that the regions, in which straight lines passing the origin has positive slopes, in the p-v
diagram are theoretically impossible, as shown in the figure. The location at which the Rayleigh
line is tangent to the reacted Hugoniot curve is the CJ plane, which was introduced previously.
The Rayleigh line intersects the unreacted Hugoniot curve at the von Neumann spike. The
pressure at the CJ plane, pCJ , can be derived from Equation B-5 by replacing p1 by pCJ :

pCJ 0 Du p p0 0 Du p

(B-14)

where p0 is approximately 0.1 MPa and can be thus ignored (the CJ pressure is approximately
21 GPa). On the reacted Hugoniot curve, the region above the CJ plane represents strong
(overdriven) detonations and that between the CJ plane and the location at v v0 weak
detonations. The overdriven detonation phenomena can occur in some special conditions such as
high velocity impacts of a flyer plate and strong detonations with properties higher than the CJ
values, thereby leading to higher detonation pressures or velocities than the CJ values. The weak

390

detonations are seldom observed but can be caused by multiple chemical reactions such as a
single reaction with positive heat release but negative volume change, two reactions of one
exothermic 14 and the other endothermic14, and transport effects, thereby attenuating energy
release or pressure (Fickett and Davis 1979), but are seldom observed. Materials that detonate at
supersonic speed are called high explosives (e.g., PETN, Nitroglycerine, RDX, TNT). The speed
of detonation is controlled by the shock wave, which causes a rapid increase in pressure and
temperature. The region below the location at p p0 on the reacted Hugoniot curve represents
deflagration, which is propagates at subsonic speed, on the order of 1-100 m/s, and the speed is
governed by heat conduction and diffusion. Deflagration is the characteristic of low explosives
(e.g., Lead azide, Mercury fulminate).

Figure B-2 States across a detonation (shock) wave in steady flow flow (adapted from
Fickett and Davis 1979)

14

In thermodynamics, the term exothermic describes a process or reaction in which the system releases energy from
its surroundings in the form of heat. On the contrary, while the term endothermic indicates a process in which the
system absorbs energy.

391

Figure B-3 Hugoniot curves and Rayleigh line (adapted from Smith and Hetherington
1994)

392

APPENDIX C
INCIDENT AND REFLECTED OVERPRESSURES VERY CLOSE TO
THE CHARGE FACE
C.1 Introduction
Numerical studies are performed using AUTODYN (ANSYS 2013a) for incident and reflected
peak overpressure very close to the face of a TNT charge with a radius of 150 mm. Two cells
sizes are selected for analysis (0.1 mm and 0.05 mm) based on the studies in Section 3.5.2. Data
are monitored at 0.33 mm intervals for distances between 149 mm and 151 mm, and at 1 mm
intervals between 147 mm and 149 mm, and between 151 mm and 155 mm.

C.2 AUTODYN Results


Results for CFD analysis are presented over a distance from 3 mm inside (147 mm, Z = 0.0517
m/kg1/3) to 5 mm beyond (155 mm, Z = 0.0545 m/kg1/3) the face of the charge. Results for the
two cell sizes are similar, as seen in Figure C-1. Inside the charge, the incident overpressures are
close to the CJ pressure (21 GPa) but increase within 1 mm of the charge face and then decrease
rapidly immediately beyond the charge face. This rapid change is attributed to the many orders
of magnitude differences in pressure, density and internal energy across the charge face. The
drop in overpressure at the charge surface is better resolved with the finer cell size of 0.05 mm.
The reflected peak overpressure at 0.33 mm from the charge surface for the 0.1 mm cells are
greater than for the 0.05 mm cells because the mesh with the 0.1 mm cells is somewhat coarse
for the distance of 0.33 mm.
The reflected overpressures at the charge face are also simulated using five small cell sizes: 1,
0.5, 0.25, 0.1, 0.05 mm. For each cell size, the reflected peak overpressure is nearly twice the
incident peak overpressure, as shown in Figure C-2, which indicates that in this case, the
influence of particle velocity is insignificant for the reflected peak overpressures. The reflected
overpressure is a combination of static pressures and dynamic (wind) pressures, where the latter
is a function of particle velocity. For the charge of TNT in direct contact with a reflecting surface,

393

Figure C-1 Incident and reflected peak overpressures near the charge face for cell sizes of
0.1 mm and 0.05 mm

Figure C-2 Incident peak overpressures inside and reflected peak overpressures at the
charge face as a function of cell size
394

shock wave reflections at the contact location occur within a solid area, before exiting the face of
the charge into the surrounding air. As particles in a solid are held at their original positions by
their intermolecular forces, as described in Appendix A, the effect of particle velocity is
insignificant in the calculation of the reflected pressure. When waves are reflected from a perfect
boundary, their amplitudes are momentarily doubled (e.g., Kolsky 1953), which is associated
with the static pressure 15 . For normal reflection, the reflected peak overpressure, pr , is
calculated by

pr 2 ps ( 1)qs

(C-1)

where ps is an incident overpressure, is the specific heat ratio, and qs is dynamic pressure.
The first and second terms correspond to the static and dynamic responses, respectively. The
particle velocity becomes very important as the shock front exits the face of the charge into the
surrounding air. Reflection coefficients for distances between 151 and 155 mm are plotted in
Figure C-3, where the reflection coefficient is the ratio of the reflected to incident overpressures.
The reflection coefficients vary between 20 and 25 at these distances, which indicates that the
reflection coefficient is related principally to particle velocity. Immediately after the shock front
exits the face of the charge, the particle acceleration is on the order of 1.0E10g. The peak particle
velocities across the charge surface for distances between 147 and 155 mm are presented in
Figure C-4. The peak particle velocity is approximately 2 km/s inside the charge face and
increases to 6 to 7 km/s at approximately 1 mm beyond the face of the charge.

15

To explain this, consider a rectangular wave having only a positive compression phase and moving
through a medium with wavelength, , velocity, v, and amplitude, A. When the wave reaches a reflecting
surface, a particle at the front of the wave exerts a force upon a particle of the reflecting surface, which
also pushes the particle at the wave front with the same force at the same time, based on the Newtons
third law. This is reflection. The reflected wave has same wavelength, , velocity, v, amplitude, A, as the
incident wave but acts in the opposite direction. The incident and reflected waves consequently merge at
the boundary during the period of / v and the amplitude of the merged waves becomes 2A before the
waves separate completely into single waves.

395

Figure C-3 Reflection coefficients for distances between 151 and 155 mm

Figure C-4 Peak particle velocities for distances between 151 and 155 mm

396

APPENDIX D
KINGERY AND BULMASH CHARTS

Figure D-1 Air-blast parameters as a function of scaled distance in SI units; spherical freeair burst of TNT
397

Figure D-2 Air-blast parameters as a function of scaled distance in US units; spherical freeair burst of TNT

398

Figure D-3 Air-blast parameters as a function of scaled distance in SI units; hemispherical


surface burst of TNT

399

Figure D-4 Air-blast parameters as a function of scaled distance in US units; hemispherical


surface burst of TNT

400

APPENDIX E
CSCM MODEL PARAMETERS IN LS-DYNA

Values of parameters of the CSCM model for concrete compressive strengths of 35.5 and 50
MPa used in LS-DYNA are presented in Table E-1. The values were chosen based on
experimental data reported by Murray (2007).
Table E-1 CSCM model parameters in LS-DYNA
Parameter

Value

Compressive strength (MPa)

35.5

50

Mass density (kg/m3)

2400

2400

Pre-existing damage

Shear modulus, G (GPa)

12.2

13.6

Bulk modulus, K (GPa)

13.3

14.9

Tri-axial compression surface constant term, (MPa)

15.2

15.9

0.3143

0.3651

10.5

10.5

Tri-axial compression surface exponent, (MPa-1)

0.01929

0.01929

Torsion surface constant term, 1

0.7473

0.7473

9.95104

4.64104

0.17

0.17

0.0646

0.04255

0.66

0.66

1.20103

556104

0.16

0.16

0.0646

0.04255

Tri-axial compression surface linear term,


Tri-axial compression surface nonlinear term, (MPa)

Torsion compression surface linear term, 1 (MPa-1)


Torsion compression surface nonlinear term, 1
Torsion compression surface exponent, 1 (MPa-1)
Tri-axial extension surface constant term, 2
Tri-axial extension surface linear term, 2 (MPa-1)
Tri-axial extension surface nonlinear term, 2
Tri-axial extension surface exponent, 2 (MPa-1)
Cap aspect ratio, R

401

Table E-1 CSCM model parameters in LS-DYNA (cont.)


Parameter

Value

Cap initial location, X0 (MPa)

93.3

104

Maximum plastic volume compaction, W

0.05

0.05

Linear shape parameter, D1 (MPa-1)

2.50104

2.50104

Quadratic shape parameter, D2 (MPa-2)

3.49107

3.49107

Ductile shape softening parameter, B

100

100

Fracture energy in uniaxial stress, Gfc (N/mm)

7.65

9.72

Brittle shape softening parameter, D

0.1

0.1

Fracture energy in uniaxial tension, Gft (N/mm)

0.0765

0.0972

Fracture energy in uniaxial tension, Gfs (N/mm)

0.0765

0.0972

Shear-to-compression transition parameter, PWRC

Shear-to-tension transition parameter, PWRT

Modify moderate pressure softening parameter, PMOD

1.90104

1.64104

Rate-effect power for uniaxial compression, Nc

0.57

0.54

Rate-effect parameter for uniaxial tension, 0t

2.1104

2.1104

Rate-effect power for uniaxial tension, Nt

0.71

0.71

Maximum stress in compression (MPa)

59.5

85

Maximum stress in tension (MPa)

59.5

85

Rate-effect parameter for uniaxial compression, 0c

402

APPENDIX F
SINGLE-DEGREE-OF-FREEDOM (SDOF) CALCULATIONS FOR
STRAIN RATE
F.1 Introduction
Single-degree-of-freedom (SDOF) calculations of strain rate are performed to validate results
from LS-DYNA for selected single element simulations presented in Section 6.5. Strain-rate
histories are generated for three linearly increasing force histories, adapted from the blast-type
loadings presented in Figure 6-13. The force histories are applied to one end of the 20 mm, 40
mm and 80 mm cubic elements. The properties of the elements are presented in Table F-1.
Table F-1 Properties of single elements
Parameter

202020 mm

404040 mm

Youngs modulus (MPa)

2.8104

Density (tonne/mm3)

2.410-9

808080 mm

Area (mm2)

400

1600

6400

Length, L (mm)

20

40

80

Volume (mm3)

8.0103

6.4104

5.1105

Stiffness, k (N/mm)

5.6105

1.1106

2.2106

Mass, m (tonne)

9.6106

7.710-5

6.110-4

Circular frequency, wn (rad/sec)

2.4105

1.2105

6.0104

Cyclic frequency, f n (Hz)

3.8104

1.9104

9.6103

Period, Tn (sec)

2.6105

5.2105

1.0104

19

14

4.8

td / Tn

403

F.2 SDOF Calculations


The equation of motion for linearly increasing force acting on a single-degree-of-freedom system
is

mu ku po

t
td

(F-1)

where m, u, k, po , t and td are the mass, displacement, stiffness, peak force, time, and time
corresponding to the peak force, respectively. The solution of Equation F-1 for 0 t td is
(Chopra 2012)

u (t )

po t sin wnt

k td
wntd

(F-2)

u (t )

po 1 cos wnt

k td
td

(F-3)

where wn is natural circular frequency.


Strain-rate histories are plotted using Equation F-3 because strain rate is proportional to velocity.
The relationship between strain rate, , and velocity, V , is
V L

(F-4)

where L is the length of an element. Figure F-1 shows that the results of the LS-DYNA analysis
and the SDOF analysis are in good agreement before the compressive strength of the concrete is
reached. The CSCM model assumes linear elastic response up to peak compressive strength, and
the strain-rate histories have similar peak amplitudes and natural periods in this range. The strain
rates computed by numerical simulations increase rapidly due to strain softening. The SDOF
analysis results validate the LS-DYNA analysis, albeit over only part of the loading phase.

404

(a) 202020 mm element, peak tensile pressure of 10 MPa, 200 MPa/msec

(b) 202020 mm element, peak tensile pressure of 20 MPa, 200 MPa/msec

(c) 404040 mm element , peak tensile pressure of 20 MPa, 50 MPa/msec


Figure F-1 Strain-rate histories for three single element simulations

405

MCEER Technical Reports


MCEER publishes technical reports on a variety of subjects written by authors funded through MCEER. These reports are
available from both MCEER Publications and the National Technical Information Service (NTIS). Requests for reports should
be directed to MCEER Publications, MCEER, University at Buffalo, State University of New York, 133A Ketter Hall, Buffalo,
New York 14260. Reports can also be requested through NTIS, P.O. Box 1425, Springfield, Virginia 22151. NTIS accession
numbers are shown in parenthesis, if available.
NCEER-87-0001

"First-Year Program in Research, Education and Technology Transfer," 3/5/87, (PB88-134275, A04, MFA01).

NCEER-87-0002

"Experimental Evaluation of Instantaneous Optimal Algorithms for Structural Control," by R.C. Lin, T.T.
Soong and A.M. Reinhorn, 4/20/87, (PB88-134341, A04, MF-A01).

NCEER-87-0003

"Experimentation Using the Earthquake Simulation Facilities at University at Buffalo," by A.M. Reinhorn
and R.L. Ketter, not available.

NCEER-87-0004

"The System Characteristics and Performance of a Shaking Table," by J.S. Hwang, K.C. Chang and G.C.
Lee, 6/1/87, (PB88-134259, A03, MF-A01). This report is available only through NTIS (see address given
above).

NCEER-87-0005

"A Finite Element Formulation for Nonlinear Viscoplastic Material Using a Q Model," by O. Gyebi and G.
Dasgupta, 11/2/87, (PB88-213764, A08, MF-A01).

NCEER-87-0006

"Symbolic Manipulation Program (SMP) - Algebraic Codes for Two and Three Dimensional Finite Element
Formulations," by X. Lee and G. Dasgupta, 11/9/87, (PB88-218522, A05, MF-A01).

NCEER-87-0007

"Instantaneous Optimal Control Laws for Tall Buildings Under Seismic Excitations," by J.N. Yang, A.
Akbarpour and P. Ghaemmaghami, 6/10/87, (PB88-134333, A06, MF-A01). This report is only available
through NTIS (see address given above).

NCEER-87-0008

"IDARC: Inelastic Damage Analysis of Reinforced Concrete Frame - Shear-Wall Structures," by Y.J. Park,
A.M. Reinhorn and S.K. Kunnath, 7/20/87, (PB88-134325, A09, MF-A01). This report is only available
through NTIS (see address given above).

NCEER-87-0009

"Liquefaction Potential for New York State: A Preliminary Report on Sites in Manhattan and Buffalo," by
M. Budhu, V. Vijayakumar, R.F. Giese and L. Baumgras, 8/31/87, (PB88-163704, A03, MF-A01). This
report is available only through NTIS (see address given above).

NCEER-87-0010

"Vertical and Torsional Vibration of Foundations in Inhomogeneous Media," by A.S. Veletsos and K.W.
Dotson, 6/1/87, (PB88-134291, A03, MF-A01). This report is only available through NTIS (see address
given above).

NCEER-87-0011

"Seismic Probabilistic Risk Assessment and Seismic Margins Studies for Nuclear Power Plants," by Howard
H.M. Hwang, 6/15/87, (PB88-134267, A03, MF-A01). This report is only available through NTIS (see
address given above).

NCEER-87-0012

"Parametric Studies of Frequency Response of Secondary Systems Under Ground-Acceleration Excitations,"


by Y. Yong and Y.K. Lin, 6/10/87, (PB88-134309, A03, MF-A01). This report is only available through
NTIS (see address given above).

NCEER-87-0013

"Frequency Response of Secondary Systems Under Seismic Excitation," by J.A. HoLung, J. Cai and Y.K.
Lin, 7/31/87, (PB88-134317, A05, MF-A01). This report is only available through NTIS (see address given
above).

NCEER-87-0014

"Modelling Earthquake Ground Motions in Seismically Active Regions Using Parametric Time Series
Methods," by G.W. Ellis and A.S. Cakmak, 8/25/87, (PB88-134283, A08, MF-A01). This report is only
available through NTIS (see address given above).

NCEER-87-0015

"Detection and Assessment of Seismic Structural Damage," by E. DiPasquale and A.S. Cakmak, 8/25/87,
(PB88-163712, A05, MF-A01). This report is only available through NTIS (see address given above).

407

NCEER-87-0016

"Pipeline Experiment at Parkfield, California," by J. Isenberg and E. Richardson, 9/15/87, (PB88-163720,


A03, MF-A01). This report is available only through NTIS (see address given above).

NCEER-87-0017

"Digital Simulation of Seismic Ground Motion," by M. Shinozuka, G. Deodatis and T. Harada, 8/31/87,
(PB88-155197, A04, MF-A01). This report is available only through NTIS (see address given above).

NCEER-87-0018

"Practical Considerations for Structural Control: System Uncertainty, System Time Delay and Truncation of
Small Control Forces," J.N. Yang and A. Akbarpour, 8/10/87, (PB88-163738, A08, MF-A01). This report is
only available through NTIS (see address given above).

NCEER-87-0019

"Modal Analysis of Nonclassically Damped Structural Systems Using Canonical Transformation," by J.N.
Yang, S. Sarkani and F.X. Long, 9/27/87, (PB88-187851, A04, MF-A01).

NCEER-87-0020

"A Nonstationary Solution in Random Vibration Theory," by J.R. Red-Horse and P.D. Spanos, 11/3/87,
(PB88-163746, A03, MF-A01).

NCEER-87-0021

"Horizontal Impedances for Radially Inhomogeneous Viscoelastic Soil Layers," by A.S. Veletsos and K.W.
Dotson, 10/15/87, (PB88-150859, A04, MF-A01).

NCEER-87-0022

"Seismic Damage Assessment of Reinforced Concrete Members," by Y.S. Chung, C. Meyer and M.
Shinozuka, 10/9/87, (PB88-150867, A05, MF-A01). This report is available only through NTIS (see address
given above).

NCEER-87-0023

"Active Structural Control in Civil Engineering," by T.T. Soong, 11/11/87, (PB88-187778, A03, MF-A01).

NCEER-87-0024

"Vertical and Torsional Impedances for Radially Inhomogeneous Viscoelastic Soil Layers," by K.W. Dotson
and A.S. Veletsos, 12/87, (PB88-187786, A03, MF-A01).

NCEER-87-0025

"Proceedings from the Symposium on Seismic Hazards, Ground Motions, Soil-Liquefaction and Engineering
Practice in Eastern North America," October 20-22, 1987, edited by K.H. Jacob, 12/87, (PB88-188115, A23,
MF-A01). This report is available only through NTIS (see address given above).

NCEER-87-0026

"Report on the Whittier-Narrows, California, Earthquake of October 1, 1987," by J. Pantelic and A.


Reinhorn, 11/87, (PB88-187752, A03, MF-A01). This report is available only through NTIS (see address
given above).

NCEER-87-0027

"Design of a Modular Program for Transient Nonlinear Analysis of Large 3-D Building Structures," by S.
Srivastav and J.F. Abel, 12/30/87, (PB88-187950, A05, MF-A01). This report is only available through NTIS
(see address given above).

NCEER-87-0028

"Second-Year Program in Research, Education and Technology Transfer," 3/8/88, (PB88-219480, A04, MFA01).

NCEER-88-0001

"Workshop on Seismic Computer Analysis and Design of Buildings With Interactive Graphics," by W.
McGuire, J.F. Abel and C.H. Conley, 1/18/88, (PB88-187760, A03, MF-A01). This report is only available
through NTIS (see address given above).

NCEER-88-0002

"Optimal Control of Nonlinear Flexible Structures," by J.N. Yang, F.X. Long and D. Wong, 1/22/88, (PB88213772, A06, MF-A01).

NCEER-88-0003

"Substructuring Techniques in the Time Domain for Primary-Secondary Structural Systems," by G.D.
Manolis and G. Juhn, 2/10/88, (PB88-213780, A04, MF-A01).

NCEER-88-0004

"Iterative Seismic Analysis of Primary-Secondary Systems," by A. Singhal, L.D. Lutes and P.D. Spanos,
2/23/88, (PB88-213798, A04, MF-A01).

NCEER-88-0005

"Stochastic Finite Element Expansion for Random Media," by P.D. Spanos and R. Ghanem, 3/14/88, (PB88213806, A03, MF-A01).

408

NCEER-88-0006

"Combining Structural Optimization and Structural Control," by F.Y. Cheng and C.P. Pantelides, 1/10/88,
(PB88-213814, A05, MF-A01).

NCEER-88-0007

"Seismic Performance Assessment of Code-Designed Structures," by H.H-M. Hwang, J-W. Jaw and H-J.
Shau, 3/20/88, (PB88-219423, A04, MF-A01). This report is only available through NTIS (see address given
above).

NCEER-88-0008

"Reliability Analysis of Code-Designed Structures Under Natural Hazards," by H.H-M. Hwang, H. Ushiba
and M. Shinozuka, 2/29/88, (PB88-229471, A07, MF-A01). This report is only available through NTIS (see
address given above).

NCEER-88-0009

"Seismic Fragility Analysis of Shear Wall Structures," by J-W Jaw and H.H-M. Hwang, 4/30/88, (PB89102867, A04, MF-A01).

NCEER-88-0010

"Base Isolation of a Multi-Story Building Under a Harmonic Ground Motion - A Comparison of


Performances of Various Systems," by F-G Fan, G. Ahmadi and I.G. Tadjbakhsh, 5/18/88, (PB89-122238,
A06, MF-A01). This report is only available through NTIS (see address given above).

NCEER-88-0011

"Seismic Floor Response Spectra for a Combined System by Green's Functions," by F.M. Lavelle, L.A.
Bergman and P.D. Spanos, 5/1/88, (PB89-102875, A03, MF-A01).

NCEER-88-0012

"A New Solution Technique for Randomly Excited Hysteretic Structures," by G.Q. Cai and Y.K. Lin,
5/16/88, (PB89-102883, A03, MF-A01).

NCEER-88-0013

"A Study of Radiation Damping and Soil-Structure Interaction Effects in the Centrifuge," by K. Weissman,
supervised by J.H. Prevost, 5/24/88, (PB89-144703, A06, MF-A01).

NCEER-88-0014

"Parameter Identification and Implementation of a Kinematic Plasticity Model for Frictional Soils," by J.H.
Prevost and D.V. Griffiths, not available.

NCEER-88-0015

"Two- and Three- Dimensional Dynamic Finite Element Analyses of the Long Valley Dam," by D.V.
Griffiths and J.H. Prevost, 6/17/88, (PB89-144711, A04, MF-A01).

NCEER-88-0016

"Damage Assessment of Reinforced Concrete Structures in Eastern United States," by A.M. Reinhorn, M.J.
Seidel, S.K. Kunnath and Y.J. Park, 6/15/88, (PB89-122220, A04, MF-A01). This report is only available
through NTIS (see address given above).

NCEER-88-0017

"Dynamic Compliance of Vertically Loaded Strip Foundations in Multilayered Viscoelastic Soils," by S.


Ahmad and A.S.M. Israil, 6/17/88, (PB89-102891, A04, MF-A01).

NCEER-88-0018

"An Experimental Study of Seismic Structural Response With Added Viscoelastic Dampers," by R.C. Lin, Z.
Liang, T.T. Soong and R.H. Zhang, 6/30/88, (PB89-122212, A05, MF-A01). This report is available only
through NTIS (see address given above).

NCEER-88-0019

"Experimental Investigation of Primary - Secondary System Interaction," by G.D. Manolis, G. Juhn and
A.M. Reinhorn, 5/27/88, (PB89-122204, A04, MF-A01).

NCEER-88-0020

"A Response Spectrum Approach For Analysis of Nonclassically Damped Structures," by J.N. Yang, S.
Sarkani and F.X. Long, 4/22/88, (PB89-102909, A04, MF-A01).

NCEER-88-0021

"Seismic Interaction of Structures and Soils: Stochastic Approach," by A.S. Veletsos and A.M. Prasad,
7/21/88, (PB89-122196, A04, MF-A01). This report is only available through NTIS (see address given
above).

NCEER-88-0022

"Identification of the Serviceability Limit State and Detection of Seismic Structural Damage," by E.
DiPasquale and A.S. Cakmak, 6/15/88, (PB89-122188, A05, MF-A01). This report is available only through
NTIS (see address given above).

NCEER-88-0023

"Multi-Hazard Risk Analysis: Case of a Simple Offshore Structure," by B.K. Bhartia and E.H. Vanmarcke,
7/21/88, (PB89-145213, A05, MF-A01).

409

NCEER-88-0024

"Automated Seismic Design of Reinforced Concrete Buildings," by Y.S. Chung, C. Meyer and M.
Shinozuka, 7/5/88, (PB89-122170, A06, MF-A01). This report is available only through NTIS (see address
given above).

NCEER-88-0025

"Experimental Study of Active Control of MDOF Structures Under Seismic Excitations," by L.L. Chung,
R.C. Lin, T.T. Soong and A.M. Reinhorn, 7/10/88, (PB89-122600, A04, MF-A01).

NCEER-88-0026

"Earthquake Simulation Tests of a Low-Rise Metal Structure," by J.S. Hwang, K.C. Chang, G.C. Lee and
R.L. Ketter, 8/1/88, (PB89-102917, A04, MF-A01).

NCEER-88-0027

"Systems Study of Urban Response and Reconstruction Due to Catastrophic Earthquakes," by F. Kozin and
H.K. Zhou, 9/22/88, (PB90-162348, A04, MF-A01).

NCEER-88-0028

"Seismic Fragility Analysis of Plane Frame Structures," by H.H-M. Hwang and Y.K. Low, 7/31/88, (PB89131445, A06, MF-A01).

NCEER-88-0029

"Response Analysis of Stochastic Structures," by A. Kardara, C. Bucher and M. Shinozuka, 9/22/88, (PB89174429, A04, MF-A01).

NCEER-88-0030

"Nonnormal Accelerations Due to Yielding in a Primary Structure," by D.C.K. Chen and L.D. Lutes,
9/19/88, (PB89-131437, A04, MF-A01).

NCEER-88-0031

"Design Approaches for Soil-Structure Interaction," by A.S. Veletsos, A.M. Prasad and Y. Tang, 12/30/88,
(PB89-174437, A03, MF-A01). This report is available only through NTIS (see address given above).

NCEER-88-0032

"A Re-evaluation of Design Spectra for Seismic Damage Control," by C.J. Turkstra and A.G. Tallin, 11/7/88,
(PB89-145221, A05, MF-A01).

NCEER-88-0033

"The Behavior and Design of Noncontact Lap Splices Subjected to Repeated Inelastic Tensile Loading," by
V.E. Sagan, P. Gergely and R.N. White, 12/8/88, (PB89-163737, A08, MF-A01).

NCEER-88-0034

"Seismic Response of Pile Foundations," by S.M. Mamoon, P.K. Banerjee and S. Ahmad, 11/1/88, (PB89145239, A04, MF-A01).

NCEER-88-0035

"Modeling of R/C Building Structures With Flexible Floor Diaphragms (IDARC2)," by A.M. Reinhorn, S.K.
Kunnath and N. Panahshahi, 9/7/88, (PB89-207153, A07, MF-A01).

NCEER-88-0036

"Solution of the Dam-Reservoir Interaction Problem Using a Combination of FEM, BEM with Particular
Integrals, Modal Analysis, and Substructuring," by C-S. Tsai, G.C. Lee and R.L. Ketter, 12/31/88, (PB89207146, A04, MF-A01).

NCEER-88-0037

"Optimal Placement of Actuators for Structural Control," by F.Y. Cheng and C.P. Pantelides, 8/15/88,
(PB89-162846, A05, MF-A01).

NCEER-88-0038

"Teflon Bearings in Aseismic Base Isolation: Experimental Studies and Mathematical Modeling," by A.
Mokha, M.C. Constantinou and A.M. Reinhorn, 12/5/88, (PB89-218457, A10, MF-A01). This report is
available only through NTIS (see address given above).

NCEER-88-0039

"Seismic Behavior of Flat Slab High-Rise Buildings in the New York City Area," by P. Weidlinger and M.
Ettouney, 10/15/88, (PB90-145681, A04, MF-A01).

NCEER-88-0040

"Evaluation of the Earthquake Resistance of Existing Buildings in New York City," by P. Weidlinger and M.
Ettouney, 10/15/88, not available.

NCEER-88-0041

"Small-Scale Modeling Techniques for Reinforced Concrete Structures Subjected to Seismic Loads," by W.
Kim, A. El-Attar and R.N. White, 11/22/88, (PB89-189625, A05, MF-A01).

NCEER-88-0042

"Modeling Strong Ground Motion from Multiple Event Earthquakes," by G.W. Ellis and A.S. Cakmak,
10/15/88, (PB89-174445, A03, MF-A01).

410

NCEER-88-0043

"Nonstationary Models of Seismic Ground Acceleration," by M. Grigoriu, S.E. Ruiz and E. Rosenblueth,
7/15/88, (PB89-189617, A04, MF-A01).

NCEER-88-0044

"SARCF User's Guide: Seismic Analysis of Reinforced Concrete Frames," by Y.S. Chung, C. Meyer and M.
Shinozuka, 11/9/88, (PB89-174452, A08, MF-A01).

NCEER-88-0045

"First Expert Panel Meeting on Disaster Research and Planning," edited by J. Pantelic and J. Stoyle, 9/15/88,
(PB89-174460, A05, MF-A01).

NCEER-88-0046

"Preliminary Studies of the Effect of Degrading Infill Walls on the Nonlinear Seismic Response of Steel
Frames," by C.Z. Chrysostomou, P. Gergely and J.F. Abel, 12/19/88, (PB89-208383, A05, MF-A01).

NCEER-88-0047

"Reinforced Concrete Frame Component Testing Facility - Design, Construction, Instrumentation and
Operation," by S.P. Pessiki, C. Conley, T. Bond, P. Gergely and R.N. White, 12/16/88, (PB89-174478, A04,
MF-A01).

NCEER-89-0001

"Effects of Protective Cushion and Soil Compliancy on the Response of Equipment Within a Seismically
Excited Building," by J.A. HoLung, 2/16/89, (PB89-207179, A04, MF-A01).

NCEER-89-0002

"Statistical Evaluation of Response Modification Factors for Reinforced Concrete Structures," by H.H-M.
Hwang and J-W. Jaw, 2/17/89, (PB89-207187, A05, MF-A01).

NCEER-89-0003

"Hysteretic Columns Under Random Excitation," by G-Q. Cai and Y.K. Lin, 1/9/89, (PB89-196513, A03,
MF-A01).

NCEER-89-0004

"Experimental Study of `Elephant Foot Bulge' Instability of Thin-Walled Metal Tanks," by Z-H. Jia and R.L.
Ketter, 2/22/89, (PB89-207195, A03, MF-A01).

NCEER-89-0005

"Experiment on Performance of Buried Pipelines Across San Andreas Fault," by J. Isenberg, E. Richardson
and T.D. O'Rourke, 3/10/89, (PB89-218440, A04, MF-A01). This report is available only through NTIS (see
address given above).

NCEER-89-0006

"A Knowledge-Based Approach to Structural Design of Earthquake-Resistant Buildings," by M. Subramani,


P. Gergely, C.H. Conley, J.F. Abel and A.H. Zaghw, 1/15/89, (PB89-218465, A06, MF-A01).

NCEER-89-0007

"Liquefaction Hazards and Their Effects on Buried Pipelines," by T.D. O'Rourke and P.A. Lane, 2/1/89,
(PB89-218481, A09, MF-A01).

NCEER-89-0008

"Fundamentals of System Identification in Structural Dynamics," by H. Imai, C-B. Yun, O. Maruyama and
M. Shinozuka, 1/26/89, (PB89-207211, A04, MF-A01).

NCEER-89-0009

"Effects of the 1985 Michoacan Earthquake on Water Systems and Other Buried Lifelines in Mexico," by
A.G. Ayala and M.J. O'Rourke, 3/8/89, (PB89-207229, A06, MF-A01).

NCEER-89-R010 "NCEER Bibliography of Earthquake Education Materials," by K.E.K. Ross, Second Revision, 9/1/89,
(PB90-125352, A05, MF-A01). This report is replaced by NCEER-92-0018.
NCEER-89-0011

"Inelastic Three-Dimensional Response Analysis of Reinforced Concrete Building Structures (IDARC-3D),


Part I - Modeling," by S.K. Kunnath and A.M. Reinhorn, 4/17/89, (PB90-114612, A07, MF-A01). This
report is available only through NTIS (see address given above).

NCEER-89-0012

"Recommended Modifications to ATC-14," by C.D. Poland and J.O. Malley, 4/12/89, (PB90-108648, A15,
MF-A01).

NCEER-89-0013

"Repair and Strengthening of Beam-to-Column Connections Subjected to Earthquake Loading," by M.


Corazao and A.J. Durrani, 2/28/89, (PB90-109885, A06, MF-A01).

NCEER-89-0014

"Program EXKAL2 for Identification of Structural Dynamic Systems," by O. Maruyama, C-B. Yun, M.
Hoshiya and M. Shinozuka, 5/19/89, (PB90-109877, A09, MF-A01).

411

NCEER-89-0015

"Response of Frames With Bolted Semi-Rigid Connections, Part I - Experimental Study and Analytical
Predictions," by P.J. DiCorso, A.M. Reinhorn, J.R. Dickerson, J.B. Radziminski and W.L. Harper, 6/1/89,
not available.

NCEER-89-0016

"ARMA Monte Carlo Simulation in Probabilistic Structural Analysis," by P.D. Spanos and M.P. Mignolet,
7/10/89, (PB90-109893, A03, MF-A01).

NCEER-89-P017

"Preliminary Proceedings from the Conference on Disaster Preparedness - The Place of Earthquake
Education in Our Schools," Edited by K.E.K. Ross, 6/23/89, (PB90-108606, A03, MF-A01).

NCEER-89-0017

"Proceedings from the Conference on Disaster Preparedness - The Place of Earthquake Education in Our
Schools," Edited by K.E.K. Ross, 12/31/89, (PB90-207895, A012, MF-A02). This report is available only
through NTIS (see address given above).

NCEER-89-0018

"Multidimensional Models of Hysteretic Material Behavior for Vibration Analysis of Shape Memory Energy
Absorbing Devices, by E.J. Graesser and F.A. Cozzarelli, 6/7/89, (PB90-164146, A04, MF-A01).

NCEER-89-0019

"Nonlinear Dynamic Analysis of Three-Dimensional Base Isolated Structures (3D-BASIS)," by S.


Nagarajaiah, A.M. Reinhorn and M.C. Constantinou, 8/3/89, (PB90-161936, A06, MF-A01). This report has
been replaced by NCEER-93-0011.

NCEER-89-0020

"Structural Control Considering Time-Rate of Control Forces and Control Rate Constraints," by F.Y. Cheng
and C.P. Pantelides, 8/3/89, (PB90-120445, A04, MF-A01).

NCEER-89-0021

"Subsurface Conditions of Memphis and Shelby County," by K.W. Ng, T-S. Chang and H-H.M. Hwang,
7/26/89, (PB90-120437, A03, MF-A01).

NCEER-89-0022

"Seismic Wave Propagation Effects on Straight Jointed Buried Pipelines," by K. Elhmadi and M.J. O'Rourke,
8/24/89, (PB90-162322, A10, MF-A02).

NCEER-89-0023

"Workshop on Serviceability Analysis of Water Delivery Systems," edited by M. Grigoriu, 3/6/89, (PB90127424, A03, MF-A01).

NCEER-89-0024

"Shaking Table Study of a 1/5 Scale Steel Frame Composed of Tapered Members," by K.C. Chang, J.S.
Hwang and G.C. Lee, 9/18/89, (PB90-160169, A04, MF-A01).

NCEER-89-0025

"DYNA1D: A Computer Program for Nonlinear Seismic Site Response Analysis - Technical
Documentation," by Jean H. Prevost, 9/14/89, (PB90-161944, A07, MF-A01). This report is available only
through NTIS (see address given above).

NCEER-89-0026

"1:4 Scale Model Studies of Active Tendon Systems and Active Mass Dampers for Aseismic Protection," by
A.M. Reinhorn, T.T. Soong, R.C. Lin, Y.P. Yang, Y. Fukao, H. Abe and M. Nakai, 9/15/89, (PB90-173246,
A10, MF-A02). This report is available only through NTIS (see address given above).

NCEER-89-0027

"Scattering of Waves by Inclusions in a Nonhomogeneous Elastic Half Space Solved by Boundary Element
Methods," by P.K. Hadley, A. Askar and A.S. Cakmak, 6/15/89, (PB90-145699, A07, MF-A01).

NCEER-89-0028

"Statistical Evaluation of Deflection Amplification Factors for Reinforced Concrete Structures," by H.H.M.
Hwang, J-W. Jaw and A.L. Ch'ng, 8/31/89, (PB90-164633, A05, MF-A01).

NCEER-89-0029

"Bedrock Accelerations in Memphis Area Due to Large New Madrid Earthquakes," by H.H.M. Hwang,
C.H.S. Chen and G. Yu, 11/7/89, (PB90-162330, A04, MF-A01).

NCEER-89-0030

"Seismic Behavior and Response Sensitivity of Secondary Structural Systems," by Y.Q. Chen and T.T.
Soong, 10/23/89, (PB90-164658, A08, MF-A01).

NCEER-89-0031

"Random Vibration and Reliability Analysis of Primary-Secondary Structural Systems," by Y. Ibrahim, M.


Grigoriu and T.T. Soong, 11/10/89, (PB90-161951, A04, MF-A01).

412

NCEER-89-0032

"Proceedings from the Second U.S. - Japan Workshop on Liquefaction, Large Ground Deformation and
Their Effects on Lifelines, September 26-29, 1989," Edited by T.D. O'Rourke and M. Hamada, 12/1/89,
(PB90-209388, A22, MF-A03).

NCEER-89-0033

"Deterministic Model for Seismic Damage Evaluation of Reinforced Concrete Structures," by J.M. Bracci,
A.M. Reinhorn, J.B. Mander and S.K. Kunnath, 9/27/89, (PB91-108803, A06, MF-A01).

NCEER-89-0034

"On the Relation Between Local and Global Damage Indices," by E. DiPasquale and A.S. Cakmak, 8/15/89,
(PB90-173865, A05, MF-A01).

NCEER-89-0035

"Cyclic Undrained Behavior of Nonplastic and Low Plasticity Silts," by A.J. Walker and H.E. Stewart,
7/26/89, (PB90-183518, A10, MF-A01).

NCEER-89-0036

"Liquefaction Potential of Surficial Deposits in the City of Buffalo, New York," by M. Budhu, R. Giese and
L. Baumgrass, 1/17/89, (PB90-208455, A04, MF-A01).

NCEER-89-0037

"A Deterministic Assessment of Effects of Ground Motion Incoherence," by A.S. Veletsos and Y. Tang,
7/15/89, (PB90-164294, A03, MF-A01).

NCEER-89-0038

"Workshop on Ground Motion Parameters for Seismic Hazard Mapping," July 17-18, 1989, edited by R.V.
Whitman, 12/1/89, (PB90-173923, A04, MF-A01).

NCEER-89-0039

"Seismic Effects on Elevated Transit Lines of the New York City Transit Authority," by C.J. Costantino,
C.A. Miller and E. Heymsfield, 12/26/89, (PB90-207887, A06, MF-A01).

NCEER-89-0040

"Centrifugal Modeling of Dynamic Soil-Structure Interaction," by K. Weissman, Supervised by J.H. Prevost,


5/10/89, (PB90-207879, A07, MF-A01).

NCEER-89-0041

"Linearized Identification of Buildings With Cores for Seismic Vulnerability Assessment," by I-K. Ho and
A.E. Aktan, 11/1/89, (PB90-251943, A07, MF-A01).

NCEER-90-0001

"Geotechnical and Lifeline Aspects of the October 17, 1989 Loma Prieta Earthquake in San Francisco," by
T.D. O'Rourke, H.E. Stewart, F.T. Blackburn and T.S. Dickerman, 1/90, (PB90-208596, A05, MF-A01).

NCEER-90-0002

"Nonnormal Secondary Response Due to Yielding in a Primary Structure," by D.C.K. Chen and L.D. Lutes,
2/28/90, (PB90-251976, A07, MF-A01).

NCEER-90-0003

"Earthquake Education Materials for Grades K-12," by K.E.K. Ross, 4/16/90, (PB91-251984, A05, MFA05). This report has been replaced by NCEER-92-0018.

NCEER-90-0004

"Catalog of Strong Motion Stations in Eastern North America," by R.W. Busby, 4/3/90, (PB90-251984, A05,
MF-A01).

NCEER-90-0005

"NCEER Strong-Motion Data Base: A User Manual for the GeoBase Release (Version 1.0 for the Sun3)," by
P. Friberg and K. Jacob, 3/31/90 (PB90-258062, A04, MF-A01).

NCEER-90-0006

"Seismic Hazard Along a Crude Oil Pipeline in the Event of an 1811-1812 Type New Madrid Earthquake,"
by H.H.M. Hwang and C-H.S. Chen, 4/16/90, (PB90-258054, A04, MF-A01).

NCEER-90-0007

"Site-Specific Response Spectra for Memphis Sheahan Pumping Station," by H.H.M. Hwang and C.S. Lee,
5/15/90, (PB91-108811, A05, MF-A01).

NCEER-90-0008

"Pilot Study on Seismic Vulnerability of Crude Oil Transmission Systems," by T. Ariman, R. Dobry, M.
Grigoriu, F. Kozin, M. O'Rourke, T. O'Rourke and M. Shinozuka, 5/25/90, (PB91-108837, A06, MF-A01).

NCEER-90-0009

"A Program to Generate Site Dependent Time Histories: EQGEN," by G.W. Ellis, M. Srinivasan and A.S.
Cakmak, 1/30/90, (PB91-108829, A04, MF-A01).

NCEER-90-0010

"Active Isolation for Seismic Protection of Operating Rooms," by M.E. Talbott, Supervised by M.
Shinozuka, 6/8/9, (PB91-110205, A05, MF-A01).

413

NCEER-90-0011

"Program LINEARID for Identification of Linear Structural Dynamic Systems," by C-B. Yun and M.
Shinozuka, 6/25/90, (PB91-110312, A08, MF-A01).

NCEER-90-0012

"Two-Dimensional Two-Phase Elasto-Plastic Seismic Response of Earth Dams," by A.N. Yiagos, Supervised
by J.H. Prevost, 6/20/90, (PB91-110197, A13, MF-A02).

NCEER-90-0013

"Secondary Systems in Base-Isolated Structures: Experimental Investigation, Stochastic Response and


Stochastic Sensitivity," by G.D. Manolis, G. Juhn, M.C. Constantinou and A.M. Reinhorn, 7/1/90, (PB91110320, A08, MF-A01).

NCEER-90-0014

"Seismic Behavior of Lightly-Reinforced Concrete Column and Beam-Column Joint Details," by S.P.
Pessiki, C.H. Conley, P. Gergely and R.N. White, 8/22/90, (PB91-108795, A11, MF-A02).

NCEER-90-0015

"Two Hybrid Control Systems for Building Structures Under Strong Earthquakes," by J.N. Yang and A.
Danielians, 6/29/90, (PB91-125393, A04, MF-A01).

NCEER-90-0016

"Instantaneous Optimal Control with Acceleration and Velocity Feedback," by J.N. Yang and Z. Li, 6/29/90,
(PB91-125401, A03, MF-A01).

NCEER-90-0017

"Reconnaissance Report on the Northern Iran Earthquake of June 21, 1990," by M. Mehrain, 10/4/90, (PB91125377, A03, MF-A01).

NCEER-90-0018

"Evaluation of Liquefaction Potential in Memphis and Shelby County," by T.S. Chang, P.S. Tang, C.S. Lee
and H. Hwang, 8/10/90, (PB91-125427, A09, MF-A01).

NCEER-90-0019

"Experimental and Analytical Study of a Combined Sliding Disc Bearing and Helical Steel Spring Isolation
System," by M.C. Constantinou, A.S. Mokha and A.M. Reinhorn, 10/4/90, (PB91-125385, A06, MF-A01).
This report is available only through NTIS (see address given above).

NCEER-90-0020

"Experimental Study and Analytical Prediction of Earthquake Response of a Sliding Isolation System with a
Spherical Surface," by A.S. Mokha, M.C. Constantinou and A.M. Reinhorn, 10/11/90, (PB91-125419, A05,
MF-A01).

NCEER-90-0021

"Dynamic Interaction Factors for Floating Pile Groups," by G. Gazetas, K. Fan, A. Kaynia and E. Kausel,
9/10/90, (PB91-170381, A05, MF-A01).

NCEER-90-0022

"Evaluation of Seismic Damage Indices for Reinforced Concrete Structures," by S. Rodriguez-Gomez and
A.S. Cakmak, 9/30/90, PB91-171322, A06, MF-A01).

NCEER-90-0023

"Study of Site Response at a Selected Memphis Site," by H. Desai, S. Ahmad, E.S. Gazetas and M.R. Oh,
10/11/90, (PB91-196857, A03, MF-A01).

NCEER-90-0024

"A User's Guide to Strongmo: Version 1.0 of NCEER's Strong-Motion Data Access Tool for PCs and
Terminals," by P.A. Friberg and C.A.T. Susch, 11/15/90, (PB91-171272, A03, MF-A01).

NCEER-90-0025

"A Three-Dimensional Analytical Study of Spatial Variability of Seismic Ground Motions," by L-L. Hong
and A.H.-S. Ang, 10/30/90, (PB91-170399, A09, MF-A01).

NCEER-90-0026

"MUMOID User's Guide - A Program for the Identification of Modal Parameters," by S. Rodriguez-Gomez
and E. DiPasquale, 9/30/90, (PB91-171298, A04, MF-A01).

NCEER-90-0027

"SARCF-II User's Guide - Seismic Analysis of Reinforced Concrete Frames," by S. Rodriguez-Gomez, Y.S.
Chung and C. Meyer, 9/30/90, (PB91-171280, A05, MF-A01).

NCEER-90-0028

"Viscous Dampers: Testing, Modeling and Application in Vibration and Seismic Isolation," by N. Makris
and M.C. Constantinou, 12/20/90 (PB91-190561, A06, MF-A01).

NCEER-90-0029

"Soil Effects on Earthquake Ground Motions in the Memphis Area," by H. Hwang, C.S. Lee, K.W. Ng and
T.S. Chang, 8/2/90, (PB91-190751, A05, MF-A01).

414

NCEER-91-0001

"Proceedings from the Third Japan-U.S. Workshop on Earthquake Resistant Design of Lifeline Facilities and
Countermeasures for Soil Liquefaction, December 17-19, 1990," edited by T.D. O'Rourke and M. Hamada,
2/1/91, (PB91-179259, A99, MF-A04).

NCEER-91-0002

"Physical Space Solutions of Non-Proportionally Damped Systems," by M. Tong, Z. Liang and G.C. Lee,
1/15/91, (PB91-179242, A04, MF-A01).

NCEER-91-0003

"Seismic Response of Single Piles and Pile Groups," by K. Fan and G. Gazetas, 1/10/91, (PB92-174994,
A04, MF-A01).

NCEER-91-0004

"Damping of Structures: Part 1 - Theory of Complex Damping," by Z. Liang and G. Lee, 10/10/91, (PB92197235, A12, MF-A03).

NCEER-91-0005

"3D-BASIS - Nonlinear Dynamic Analysis of Three Dimensional Base Isolated Structures: Part II," by S.
Nagarajaiah, A.M. Reinhorn and M.C. Constantinou, 2/28/91, (PB91-190553, A07, MF-A01). This report
has been replaced by NCEER-93-0011.

NCEER-91-0006

"A Multidimensional Hysteretic Model for Plasticity Deforming Metals in Energy Absorbing Devices," by
E.J. Graesser and F.A. Cozzarelli, 4/9/91, (PB92-108364, A04, MF-A01).

NCEER-91-0007

"A Framework for Customizable Knowledge-Based Expert Systems with an Application to a KBES for
Evaluating the Seismic Resistance of Existing Buildings," by E.G. Ibarra-Anaya and S.J. Fenves, 4/9/91,
(PB91-210930, A08, MF-A01).

NCEER-91-0008

"Nonlinear Analysis of Steel Frames with Semi-Rigid Connections Using the Capacity Spectrum Method,"
by G.G. Deierlein, S-H. Hsieh, Y-J. Shen and J.F. Abel, 7/2/91, (PB92-113828, A05, MF-A01).

NCEER-91-0009

"Earthquake Education Materials for Grades K-12," by K.E.K. Ross, 4/30/91, (PB91-212142, A06, MFA01). This report has been replaced by NCEER-92-0018.

NCEER-91-0010

"Phase Wave Velocities and Displacement Phase Differences in a Harmonically Oscillating Pile," by N.
Makris and G. Gazetas, 7/8/91, (PB92-108356, A04, MF-A01).

NCEER-91-0011

"Dynamic Characteristics of a Full-Size Five-Story Steel Structure and a 2/5 Scale Model," by K.C. Chang,
G.C. Yao, G.C. Lee, D.S. Hao and Y.C. Yeh," 7/2/91, (PB93-116648, A06, MF-A02).

NCEER-91-0012

"Seismic Response of a 2/5 Scale Steel Structure with Added Viscoelastic Dampers," by K.C. Chang, T.T.
Soong, S-T. Oh and M.L. Lai, 5/17/91, (PB92-110816, A05, MF-A01).

NCEER-91-0013

"Earthquake Response of Retaining Walls; Full-Scale Testing and Computational Modeling," by S.


Alampalli and A-W.M. Elgamal, 6/20/91, not available.

NCEER-91-0014

"3D-BASIS-M: Nonlinear Dynamic Analysis of Multiple Building Base Isolated Structures," by P.C.
Tsopelas, S. Nagarajaiah, M.C. Constantinou and A.M. Reinhorn, 5/28/91, (PB92-113885, A09, MF-A02).

NCEER-91-0015

"Evaluation of SEAOC Design Requirements for Sliding Isolated Structures," by D. Theodossiou and M.C.
Constantinou, 6/10/91, (PB92-114602, A11, MF-A03).

NCEER-91-0016

"Closed-Loop Modal Testing of a 27-Story Reinforced Concrete Flat Plate-Core Building," by H.R.
Somaprasad, T. Toksoy, H. Yoshiyuki and A.E. Aktan, 7/15/91, (PB92-129980, A07, MF-A02).

NCEER-91-0017

"Shake Table Test of a 1/6 Scale Two-Story Lightly Reinforced Concrete Building," by A.G. El-Attar, R.N.
White and P. Gergely, 2/28/91, (PB92-222447, A06, MF-A02).

NCEER-91-0018

"Shake Table Test of a 1/8 Scale Three-Story Lightly Reinforced Concrete Building," by A.G. El-Attar, R.N.
White and P. Gergely, 2/28/91, (PB93-116630, A08, MF-A02).

NCEER-91-0019

"Transfer Functions for Rigid Rectangular Foundations," by A.S. Veletsos, A.M. Prasad and W.H. Wu,
7/31/91, not available.

415

NCEER-91-0020

"Hybrid Control of Seismic-Excited Nonlinear and Inelastic Structural Systems," by J.N. Yang, Z. Li and A.
Danielians, 8/1/91, (PB92-143171, A06, MF-A02).

NCEER-91-0021

"The NCEER-91 Earthquake Catalog: Improved Intensity-Based Magnitudes and Recurrence Relations for
U.S. Earthquakes East of New Madrid," by L. Seeber and J.G. Armbruster, 8/28/91, (PB92-176742, A06,
MF-A02).

NCEER-91-0022

"Proceedings from the Implementation of Earthquake Planning and Education in Schools: The Need for
Change - The Roles of the Changemakers," by K.E.K. Ross and F. Winslow, 7/23/91, (PB92-129998, A12,
MF-A03).

NCEER-91-0023

"A Study of Reliability-Based Criteria for Seismic Design of Reinforced Concrete Frame Buildings," by
H.H.M. Hwang and H-M. Hsu, 8/10/91, (PB92-140235, A09, MF-A02).

NCEER-91-0024

"Experimental Verification of a Number of Structural System Identification Algorithms," by R.G. Ghanem,


H. Gavin and M. Shinozuka, 9/18/91, (PB92-176577, A18, MF-A04).

NCEER-91-0025

"Probabilistic Evaluation of Liquefaction Potential," by H.H.M. Hwang and C.S. Lee," 11/25/91, (PB92143429, A05, MF-A01).

NCEER-91-0026

"Instantaneous Optimal Control for Linear, Nonlinear and Hysteretic Structures - Stable Controllers," by J.N.
Yang and Z. Li, 11/15/91, (PB92-163807, A04, MF-A01).

NCEER-91-0027

"Experimental and Theoretical Study of a Sliding Isolation System for Bridges," by M.C. Constantinou, A.
Kartoum, A.M. Reinhorn and P. Bradford, 11/15/91, (PB92-176973, A10, MF-A03).

NCEER-92-0001

"Case Studies of Liquefaction and Lifeline Performance During Past Earthquakes, Volume 1: Japanese Case
Studies," Edited by M. Hamada and T. O'Rourke, 2/17/92, (PB92-197243, A18, MF-A04).

NCEER-92-0002

"Case Studies of Liquefaction and Lifeline Performance During Past Earthquakes, Volume 2: United States
Case Studies," Edited by T. O'Rourke and M. Hamada, 2/17/92, (PB92-197250, A20, MF-A04).

NCEER-92-0003

"Issues in Earthquake Education," Edited by K. Ross, 2/3/92, (PB92-222389, A07, MF-A02).

NCEER-92-0004

"Proceedings from the First U.S. - Japan Workshop on Earthquake Protective Systems for Bridges," Edited
by I.G. Buckle, 2/4/92, (PB94-142239, A99, MF-A06).

NCEER-92-0005

"Seismic Ground Motion from a Haskell-Type Source in a Multiple-Layered Half-Space," A.P. Theoharis, G.
Deodatis and M. Shinozuka, 1/2/92, not available.

NCEER-92-0006

"Proceedings from the Site Effects Workshop," Edited by R. Whitman, 2/29/92, (PB92-197201, A04, MFA01).

NCEER-92-0007

"Engineering Evaluation of Permanent Ground Deformations Due to Seismically-Induced Liquefaction," by


M.H. Baziar, R. Dobry and A-W.M. Elgamal, 3/24/92, (PB92-222421, A13, MF-A03).

NCEER-92-0008

"A Procedure for the Seismic Evaluation of Buildings in the Central and Eastern United States," by C.D.
Poland and J.O. Malley, 4/2/92, (PB92-222439, A20, MF-A04).

NCEER-92-0009

"Experimental and Analytical Study of a Hybrid Isolation System Using Friction Controllable Sliding
Bearings," by M.Q. Feng, S. Fujii and M. Shinozuka, 5/15/92, (PB93-150282, A06, MF-A02).

NCEER-92-0010

"Seismic Resistance of Slab-Column Connections in Existing Non-Ductile Flat-Plate Buildings," by A.J.


Durrani and Y. Du, 5/18/92, (PB93-116812, A06, MF-A02).

NCEER-92-0011

"The Hysteretic and Dynamic Behavior of Brick Masonry Walls Upgraded by Ferrocement Coatings Under
Cyclic Loading and Strong Simulated Ground Motion," by H. Lee and S.P. Prawel, 5/11/92, not available.

NCEER-92-0012

"Study of Wire Rope Systems for Seismic Protection of Equipment in Buildings," by G.F. Demetriades,
M.C. Constantinou and A.M. Reinhorn, 5/20/92, (PB93-116655, A08, MF-A02).

416

NCEER-92-0013

"Shape Memory Structural Dampers: Material Properties, Design and Seismic Testing," by P.R. Witting and
F.A. Cozzarelli, 5/26/92, (PB93-116663, A05, MF-A01).

NCEER-92-0014

"Longitudinal Permanent Ground Deformation Effects on Buried Continuous Pipelines," by M.J. O'Rourke,
and C. Nordberg, 6/15/92, (PB93-116671, A08, MF-A02).

NCEER-92-0015

"A Simulation Method for Stationary Gaussian Random Functions Based on the Sampling Theorem," by M.
Grigoriu and S. Balopoulou, 6/11/92, (PB93-127496, A05, MF-A01).

NCEER-92-0016

"Gravity-Load-Designed Reinforced Concrete Buildings: Seismic Evaluation of Existing Construction and


Detailing Strategies for Improved Seismic Resistance," by G.W. Hoffmann, S.K. Kunnath, A.M. Reinhorn
and J.B. Mander, 7/15/92, (PB94-142007, A08, MF-A02).

NCEER-92-0017

"Observations on Water System and Pipeline Performance in the Limn Area of Costa Rica Due to the April
22, 1991 Earthquake," by M. O'Rourke and D. Ballantyne, 6/30/92, (PB93-126811, A06, MF-A02).

NCEER-92-0018

"Fourth Edition of Earthquake Education Materials for Grades K-12," Edited by K.E.K. Ross, 8/10/92,
(PB93-114023, A07, MF-A02).

NCEER-92-0019

"Proceedings from the Fourth Japan-U.S. Workshop on Earthquake Resistant Design of Lifeline Facilities
and Countermeasures for Soil Liquefaction," Edited by M. Hamada and T.D. O'Rourke, 8/12/92, (PB93163939, A99, MF-E11).

NCEER-92-0020

"Active Bracing System: A Full Scale Implementation of Active Control," by A.M. Reinhorn, T.T. Soong,
R.C. Lin, M.A. Riley, Y.P. Wang, S. Aizawa and M. Higashino, 8/14/92, (PB93-127512, A06, MF-A02).

NCEER-92-0021

"Empirical Analysis of Horizontal Ground Displacement Generated by Liquefaction-Induced Lateral


Spreads," by S.F. Bartlett and T.L. Youd, 8/17/92, (PB93-188241, A06, MF-A02).

NCEER-92-0022

"IDARC Version 3.0: Inelastic Damage Analysis of Reinforced Concrete Structures," by S.K. Kunnath, A.M.
Reinhorn and R.F. Lobo, 8/31/92, (PB93-227502, A07, MF-A02).

NCEER-92-0023

"A Semi-Empirical Analysis of Strong-Motion Peaks in Terms of Seismic Source, Propagation Path and
Local Site Conditions, by M. Kamiyama, M.J. O'Rourke and R. Flores-Berrones, 9/9/92, (PB93-150266,
A08, MF-A02).

NCEER-92-0024

"Seismic Behavior of Reinforced Concrete Frame Structures with Nonductile Details, Part I: Summary of
Experimental Findings of Full Scale Beam-Column Joint Tests," by A. Beres, R.N. White and P. Gergely,
9/30/92, (PB93-227783, A05, MF-A01).

NCEER-92-0025

"Experimental Results of Repaired and Retrofitted Beam-Column Joint Tests in Lightly Reinforced Concrete
Frame Buildings," by A. Beres, S. El-Borgi, R.N. White and P. Gergely, 10/29/92, (PB93-227791, A05, MFA01).

NCEER-92-0026

"A Generalization of Optimal Control Theory: Linear and Nonlinear Structures," by J.N. Yang, Z. Li and S.
Vongchavalitkul, 11/2/92, (PB93-188621, A05, MF-A01).

NCEER-92-0027

"Seismic Resistance of Reinforced Concrete Frame Structures Designed Only for Gravity Loads: Part I Design and Properties of a One-Third Scale Model Structure," by J.M. Bracci, A.M. Reinhorn and J.B.
Mander, 12/1/92, (PB94-104502, A08, MF-A02).

NCEER-92-0028

"Seismic Resistance of Reinforced Concrete Frame Structures Designed Only for Gravity Loads: Part II Experimental Performance of Subassemblages," by L.E. Aycardi, J.B. Mander and A.M. Reinhorn, 12/1/92,
(PB94-104510, A08, MF-A02).

NCEER-92-0029

"Seismic Resistance of Reinforced Concrete Frame Structures Designed Only for Gravity Loads: Part III Experimental Performance and Analytical Study of a Structural Model," by J.M. Bracci, A.M. Reinhorn and
J.B. Mander, 12/1/92, (PB93-227528, A09, MF-A01).

417

NCEER-92-0030

"Evaluation of Seismic Retrofit of Reinforced Concrete Frame Structures: Part I - Experimental Performance
of Retrofitted Subassemblages," by D. Choudhuri, J.B. Mander and A.M. Reinhorn, 12/8/92, (PB93-198307,
A07, MF-A02).

NCEER-92-0031

"Evaluation of Seismic Retrofit of Reinforced Concrete Frame Structures: Part II - Experimental


Performance and Analytical Study of a Retrofitted Structural Model," by J.M. Bracci, A.M. Reinhorn and
J.B. Mander, 12/8/92, (PB93-198315, A09, MF-A03).

NCEER-92-0032

"Experimental and Analytical Investigation of Seismic Response of Structures with Supplemental Fluid
Viscous Dampers," by M.C. Constantinou and M.D. Symans, 12/21/92, (PB93-191435, A10, MF-A03). This
report is available only through NTIS (see address given above).

NCEER-92-0033

"Reconnaissance Report on the Cairo, Egypt Earthquake of October 12, 1992," by M. Khater, 12/23/92,
(PB93-188621, A03, MF-A01).

NCEER-92-0034

"Low-Level Dynamic Characteristics of Four Tall Flat-Plate Buildings in New York City," by H. Gavin, S.
Yuan, J. Grossman, E. Pekelis and K. Jacob, 12/28/92, (PB93-188217, A07, MF-A02).

NCEER-93-0001

"An Experimental Study on the Seismic Performance of Brick-Infilled Steel Frames With and Without
Retrofit," by J.B. Mander, B. Nair, K. Wojtkowski and J. Ma, 1/29/93, (PB93-227510, A07, MF-A02).

NCEER-93-0002

"Social Accounting for Disaster Preparedness and Recovery Planning," by S. Cole, E. Pantoja and V. Razak,
2/22/93, (PB94-142114, A12, MF-A03).

NCEER-93-0003

"Assessment of 1991 NEHRP Provisions for Nonstructural Components and Recommended Revisions," by
T.T. Soong, G. Chen, Z. Wu, R-H. Zhang and M. Grigoriu, 3/1/93, (PB93-188639, A06, MF-A02).

NCEER-93-0004

"Evaluation of Static and Response Spectrum Analysis Procedures of SEAOC/UBC for Seismic Isolated
Structures," by C.W. Winters and M.C. Constantinou, 3/23/93, (PB93-198299, A10, MF-A03).

NCEER-93-0005

"Earthquakes in the Northeast - Are We Ignoring the Hazard? A Workshop on Earthquake Science and
Safety for Educators," edited by K.E.K. Ross, 4/2/93, (PB94-103066, A09, MF-A02).

NCEER-93-0006

"Inelastic Response of Reinforced Concrete Structures with Viscoelastic Braces," by R.F. Lobo, J.M. Bracci,
K.L. Shen, A.M. Reinhorn and T.T. Soong, 4/5/93, (PB93-227486, A05, MF-A02).

NCEER-93-0007

"Seismic Testing of Installation Methods for Computers and Data Processing Equipment," by K. Kosar, T.T.
Soong, K.L. Shen, J.A. HoLung and Y.K. Lin, 4/12/93, (PB93-198299, A07, MF-A02).

NCEER-93-0008

"Retrofit of Reinforced Concrete Frames Using Added Dampers," by A. Reinhorn, M. Constantinou and C.
Li, not available.

NCEER-93-0009

"Seismic Behavior and Design Guidelines for Steel Frame Structures with Added Viscoelastic Dampers," by
K.C. Chang, M.L. Lai, T.T. Soong, D.S. Hao and Y.C. Yeh, 5/1/93, (PB94-141959, A07, MF-A02).

NCEER-93-0010

"Seismic Performance of Shear-Critical Reinforced Concrete Bridge Piers," by J.B. Mander, S.M. Waheed,
M.T.A. Chaudhary and S.S. Chen, 5/12/93, (PB93-227494, A08, MF-A02).

NCEER-93-0011

"3D-BASIS-TABS: Computer Program for Nonlinear Dynamic Analysis of Three Dimensional Base Isolated
Structures," by S. Nagarajaiah, C. Li, A.M. Reinhorn and M.C. Constantinou, 8/2/93, (PB94-141819, A09,
MF-A02).

NCEER-93-0012

"Effects of Hydrocarbon Spills from an Oil Pipeline Break on Ground Water," by O.J. Helweg and H.H.M.
Hwang, 8/3/93, (PB94-141942, A06, MF-A02).

NCEER-93-0013

"Simplified Procedures for Seismic Design of Nonstructural Components and Assessment of Current Code
Provisions," by M.P. Singh, L.E. Suarez, E.E. Matheu and G.O. Maldonado, 8/4/93, (PB94-141827, A09,
MF-A02).

NCEER-93-0014

"An Energy Approach to Seismic Analysis and Design of Secondary Systems," by G. Chen and T.T. Soong,
8/6/93, (PB94-142767, A11, MF-A03).

418

NCEER-93-0015

"Proceedings from School Sites: Becoming Prepared for Earthquakes - Commemorating the Third
Anniversary of the Loma Prieta Earthquake," Edited by F.E. Winslow and K.E.K. Ross, 8/16/93, (PB94154275, A16, MF-A02).

NCEER-93-0016

"Reconnaissance Report of Damage to Historic Monuments in Cairo, Egypt Following the October 12, 1992
Dahshur Earthquake," by D. Sykora, D. Look, G. Croci, E. Karaesmen and E. Karaesmen, 8/19/93, (PB94142221, A08, MF-A02).

NCEER-93-0017

"The Island of Guam Earthquake of August 8, 1993," by S.W. Swan and S.K. Harris, 9/30/93, (PB94141843, A04, MF-A01).

NCEER-93-0018

"Engineering Aspects of the October 12, 1992 Egyptian Earthquake," by A.W. Elgamal, M. Amer, K.
Adalier and A. Abul-Fadl, 10/7/93, (PB94-141983, A05, MF-A01).

NCEER-93-0019

"Development of an Earthquake Motion Simulator and its Application in Dynamic Centrifuge Testing," by I.
Krstelj, Supervised by J.H. Prevost, 10/23/93, (PB94-181773, A-10, MF-A03).

NCEER-93-0020

"NCEER-Taisei Corporation Research Program on Sliding Seismic Isolation Systems for Bridges:
Experimental and Analytical Study of a Friction Pendulum System (FPS)," by M.C. Constantinou, P.
Tsopelas, Y-S. Kim and S. Okamoto, 11/1/93, (PB94-142775, A08, MF-A02).

NCEER-93-0021

"Finite Element Modeling of Elastomeric Seismic Isolation Bearings," by L.J. Billings, Supervised by R.
Shepherd, 11/8/93, not available.

NCEER-93-0022

"Seismic Vulnerability of Equipment in Critical Facilities: Life-Safety and Operational Consequences," by


K. Porter, G.S. Johnson, M.M. Zadeh, C. Scawthorn and S. Eder, 11/24/93, (PB94-181765, A16, MF-A03).

NCEER-93-0023

"Hokkaido Nansei-oki, Japan Earthquake of July 12, 1993, by P.I. Yanev and C.R. Scawthorn, 12/23/93,
(PB94-181500, A07, MF-A01).

NCEER-94-0001

"An Evaluation of Seismic Serviceability of Water Supply Networks with Application to the San Francisco
Auxiliary Water Supply System," by I. Markov, Supervised by M. Grigoriu and T. O'Rourke, 1/21/94,
(PB94-204013, A07, MF-A02).

NCEER-94-0002

"NCEER-Taisei Corporation Research Program on Sliding Seismic Isolation Systems for Bridges:
Experimental and Analytical Study of Systems Consisting of Sliding Bearings, Rubber Restoring Force
Devices and Fluid Dampers," Volumes I and II, by P. Tsopelas, S. Okamoto, M.C. Constantinou, D. Ozaki
and S. Fujii, 2/4/94, (PB94-181740, A09, MF-A02 and PB94-181757, A12, MF-A03).

NCEER-94-0003

"A Markov Model for Local and Global Damage Indices in Seismic Analysis," by S. Rahman and M.
Grigoriu, 2/18/94, (PB94-206000, A12, MF-A03).

NCEER-94-0004

"Proceedings from the NCEER Workshop on Seismic Response of Masonry Infills," edited by D.P. Abrams,
3/1/94, (PB94-180783, A07, MF-A02).

NCEER-94-0005

"The Northridge, California Earthquake of January 17, 1994: General Reconnaissance Report," edited by
J.D. Goltz, 3/11/94, (PB94-193943, A10, MF-A03).

NCEER-94-0006

"Seismic Energy Based Fatigue Damage Analysis of Bridge Columns: Part I - Evaluation of Seismic
Capacity," by G.A. Chang and J.B. Mander, 3/14/94, (PB94-219185, A11, MF-A03).

NCEER-94-0007

"Seismic Isolation of Multi-Story Frame Structures Using Spherical Sliding Isolation Systems," by T.M. AlHussaini, V.A. Zayas and M.C. Constantinou, 3/17/94, (PB94-193745, A09, MF-A02).

NCEER-94-0008

"The Northridge, California Earthquake of January 17, 1994: Performance of Highway Bridges," edited by
I.G. Buckle, 3/24/94, (PB94-193851, A06, MF-A02).

NCEER-94-0009

"Proceedings of the Third U.S.-Japan Workshop on Earthquake Protective Systems for Bridges," edited by
I.G. Buckle and I. Friedland, 3/31/94, (PB94-195815, A99, MF-A06).

419

NCEER-94-0010

"3D-BASIS-ME: Computer Program for Nonlinear Dynamic Analysis of Seismically Isolated Single and
Multiple Structures and Liquid Storage Tanks," by P.C. Tsopelas, M.C. Constantinou and A.M. Reinhorn,
4/12/94, (PB94-204922, A09, MF-A02).

NCEER-94-0011

"The Northridge, California Earthquake of January 17, 1994: Performance of Gas Transmission Pipelines,"
by T.D. O'Rourke and M.C. Palmer, 5/16/94, (PB94-204989, A05, MF-A01).

NCEER-94-0012

"Feasibility Study of Replacement Procedures and Earthquake Performance Related to Gas Transmission
Pipelines," by T.D. O'Rourke and M.C. Palmer, 5/25/94, (PB94-206638, A09, MF-A02).

NCEER-94-0013

"Seismic Energy Based Fatigue Damage Analysis of Bridge Columns: Part II - Evaluation of Seismic
Demand," by G.A. Chang and J.B. Mander, 6/1/94, (PB95-18106, A08, MF-A02).

NCEER-94-0014

"NCEER-Taisei Corporation Research Program on Sliding Seismic Isolation Systems for Bridges:
Experimental and Analytical Study of a System Consisting of Sliding Bearings and Fluid Restoring
Force/Damping Devices," by P. Tsopelas and M.C. Constantinou, 6/13/94, (PB94-219144, A10, MF-A03).

NCEER-94-0015

"Generation of Hazard-Consistent Fragility Curves for Seismic Loss Estimation Studies," by H. Hwang and
J-R. Huo, 6/14/94, (PB95-181996, A09, MF-A02).

NCEER-94-0016

"Seismic Study of Building Frames with Added Energy-Absorbing Devices," by W.S. Pong, C.S. Tsai and
G.C. Lee, 6/20/94, (PB94-219136, A10, A03).

NCEER-94-0017

"Sliding Mode Control for Seismic-Excited Linear and Nonlinear Civil Engineering Structures," by J. Yang,
J. Wu, A. Agrawal and Z. Li, 6/21/94, (PB95-138483, A06, MF-A02).

NCEER-94-0018

"3D-BASIS-TABS Version 2.0: Computer Program for Nonlinear Dynamic Analysis of Three Dimensional
Base Isolated Structures," by A.M. Reinhorn, S. Nagarajaiah, M.C. Constantinou, P. Tsopelas and R. Li,
6/22/94, (PB95-182176, A08, MF-A02).

NCEER-94-0019

"Proceedings of the International Workshop on Civil Infrastructure Systems: Application of Intelligent


Systems and Advanced Materials on Bridge Systems," Edited by G.C. Lee and K.C. Chang, 7/18/94, (PB95252474, A20, MF-A04).

NCEER-94-0020

"Study of Seismic Isolation Systems for Computer Floors," by V. Lambrou and M.C. Constantinou, 7/19/94,
(PB95-138533, A10, MF-A03).

NCEER-94-0021

"Proceedings of the U.S.-Italian Workshop on Guidelines for Seismic Evaluation and Rehabilitation of
Unreinforced Masonry Buildings," Edited by D.P. Abrams and G.M. Calvi, 7/20/94, (PB95-138749, A13,
MF-A03).

NCEER-94-0022

"NCEER-Taisei Corporation Research Program on Sliding Seismic Isolation Systems for Bridges:
Experimental and Analytical Study of a System Consisting of Lubricated PTFE Sliding Bearings and Mild
Steel Dampers," by P. Tsopelas and M.C. Constantinou, 7/22/94, (PB95-182184, A08, MF-A02).

NCEER-94-0023

Development of Reliability-Based Design Criteria for Buildings Under Seismic Load, by Y.K. Wen, H.
Hwang and M. Shinozuka, 8/1/94, (PB95-211934, A08, MF-A02).

NCEER-94-0024

Experimental Verification of Acceleration Feedback Control Strategies for an Active Tendon System, by
S.J. Dyke, B.F. Spencer, Jr., P. Quast, M.K. Sain, D.C. Kaspari, Jr. and T.T. Soong, 8/29/94, (PB95-212320,
A05, MF-A01).

NCEER-94-0025

Seismic Retrofitting Manual for Highway Bridges, Edited by I.G. Buckle and I.F. Friedland, published by
the Federal Highway Administration (PB95-212676, A15, MF-A03).

NCEER-94-0026

Proceedings from the Fifth U.S.-Japan Workshop on Earthquake Resistant Design of Lifeline Facilities and
Countermeasures Against Soil Liquefaction, Edited by T.D. ORourke and M. Hamada, 11/7/94, (PB95220802, A99, MF-E08).

420

NCEER-95-0001

Experimental and Analytical Investigation of Seismic Retrofit of Structures with Supplemental Damping:
Part 1 - Fluid Viscous Damping Devices, by A.M. Reinhorn, C. Li and M.C. Constantinou, 1/3/95, (PB95266599, A09, MF-A02).

NCEER-95-0002

Experimental and Analytical Study of Low-Cycle Fatigue Behavior of Semi-Rigid Top-And-Seat Angle
Connections, by G. Pekcan, J.B. Mander and S.S. Chen, 1/5/95, (PB95-220042, A07, MF-A02).

NCEER-95-0003

NCEER-ATC Joint Study on Fragility of Buildings, by T. Anagnos, C. Rojahn and A.S. Kiremidjian,
1/20/95, (PB95-220026, A06, MF-A02).

NCEER-95-0004

Nonlinear Control Algorithms for Peak Response Reduction, by Z. Wu, T.T. Soong, V. Gattulli and R.C.
Lin, 2/16/95, (PB95-220349, A05, MF-A01).

NCEER-95-0005

Pipeline Replacement Feasibility Study: A Methodology for Minimizing Seismic and Corrosion Risks to
Underground Natural Gas Pipelines, by R.T. Eguchi, H.A. Seligson and D.G. Honegger, 3/2/95, (PB95252326, A06, MF-A02).

NCEER-95-0006

Evaluation of Seismic Performance of an 11-Story Frame Building During the 1994 Northridge
Earthquake, by F. Naeim, R. DiSulio, K. Benuska, A. Reinhorn and C. Li, not available.

NCEER-95-0007

Prioritization of Bridges for Seismic Retrofitting, by N. Basz and A.S. Kiremidjian, 4/24/95, (PB95252300, A08, MF-A02).

NCEER-95-0008

Method for Developing Motion Damage Relationships for Reinforced Concrete Frames, by A. Singhal and
A.S. Kiremidjian, 5/11/95, (PB95-266607, A06, MF-A02).

NCEER-95-0009

Experimental and Analytical Investigation of Seismic Retrofit of Structures with Supplemental Damping:
Part II - Friction Devices, by C. Li and A.M. Reinhorn, 7/6/95, (PB96-128087, A11, MF-A03).

NCEER-95-0010

Experimental Performance and Analytical Study of a Non-Ductile Reinforced Concrete Frame Structure
Retrofitted with Elastomeric Spring Dampers, by G. Pekcan, J.B. Mander and S.S. Chen, 7/14/95, (PB96137161, A08, MF-A02).

NCEER-95-0011

Development and Experimental Study of Semi-Active Fluid Damping Devices for Seismic Protection of
Structures, by M.D. Symans and M.C. Constantinou, 8/3/95, (PB96-136940, A23, MF-A04).

NCEER-95-0012

Real-Time Structural Parameter Modification (RSPM): Development of Innervated Structures, by Z.


Liang, M. Tong and G.C. Lee, 4/11/95, (PB96-137153, A06, MF-A01).

NCEER-95-0013

Experimental and Analytical Investigation of Seismic Retrofit of Structures with Supplemental Damping:
Part III - Viscous Damping Walls, by A.M. Reinhorn and C. Li, 10/1/95, (PB96-176409, A11, MF-A03).

NCEER-95-0014

Seismic Fragility Analysis of Equipment and Structures in a Memphis Electric Substation, by J-R. Huo and
H.H.M. Hwang, 8/10/95, (PB96-128087, A09, MF-A02).

NCEER-95-0015

The Hanshin-Awaji Earthquake of January 17, 1995: Performance of Lifelines, Edited by M. Shinozuka,
11/3/95, (PB96-176383, A15, MF-A03).

NCEER-95-0016

Highway Culvert Performance During Earthquakes, by T.L. Youd and C.J. Beckman, available as
NCEER-96-0015.

NCEER-95-0017

The Hanshin-Awaji Earthquake of January 17, 1995: Performance of Highway Bridges, Edited by I.G.
Buckle, 12/1/95, not available.

NCEER-95-0018

Modeling of Masonry Infill Panels for Structural Analysis, by A.M. Reinhorn, A. Madan, R.E. Valles, Y.
Reichmann and J.B. Mander, 12/8/95, (PB97-110886, MF-A01, A06).

NCEER-95-0019

Optimal Polynomial Control for Linear and Nonlinear Structures, by A.K. Agrawal and J.N. Yang,
12/11/95, (PB96-168737, A07, MF-A02).

421

NCEER-95-0020

Retrofit of Non-Ductile Reinforced Concrete Frames Using Friction Dampers, by R.S. Rao, P. Gergely and
R.N. White, 12/22/95, (PB97-133508, A10, MF-A02).

NCEER-95-0021

Parametric Results for Seismic Response of Pile-Supported Bridge Bents, by G. Mylonakis, A. Nikolaou
and G. Gazetas, 12/22/95, (PB97-100242, A12, MF-A03).

NCEER-95-0022

Kinematic Bending Moments in Seismically Stressed Piles, by A. Nikolaou, G. Mylonakis and G. Gazetas,
12/23/95, (PB97-113914, MF-A03, A13).

NCEER-96-0001

Dynamic Response of Unreinforced Masonry Buildings with Flexible Diaphragms, by A.C. Costley and
D.P. Abrams, 10/10/96, (PB97-133573, MF-A03, A15).

NCEER-96-0002

State of the Art Review: Foundations and Retaining Structures, by I. Po Lam, not available.

NCEER-96-0003

Ductility of Rectangular Reinforced Concrete Bridge Columns with Moderate Confinement, by N. Wehbe,
M. Saiidi, D. Sanders and B. Douglas, 11/7/96, (PB97-133557, A06, MF-A02).

NCEER-96-0004

Proceedings of the Long-Span Bridge Seismic Research Workshop, edited by I.G. Buckle and I.M.
Friedland, not available.

NCEER-96-0005

Establish Representative Pier Types for Comprehensive Study: Eastern United States, by J. Kulicki and Z.
Prucz, 5/28/96, (PB98-119217, A07, MF-A02).

NCEER-96-0006

Establish Representative Pier Types for Comprehensive Study: Western United States, by R. Imbsen, R.A.
Schamber and T.A. Osterkamp, 5/28/96, (PB98-118607, A07, MF-A02).

NCEER-96-0007

Nonlinear Control Techniques for Dynamical Systems with Uncertain Parameters, by R.G. Ghanem and
M.I. Bujakov, 5/27/96, (PB97-100259, A17, MF-A03).

NCEER-96-0008

Seismic Evaluation of a 30-Year Old Non-Ductile Highway Bridge Pier and Its Retrofit, by J.B. Mander,
B. Mahmoodzadegan, S. Bhadra and S.S. Chen, 5/31/96, (PB97-110902, MF-A03, A10).

NCEER-96-0009

Seismic Performance of a Model Reinforced Concrete Bridge Pier Before and After Retrofit, by J.B.
Mander, J.H. Kim and C.A. Ligozio, 5/31/96, (PB97-110910, MF-A02, A10).

NCEER-96-0010

IDARC2D Version 4.0: A Computer Program for the Inelastic Damage Analysis of Buildings, by R.E.
Valles, A.M. Reinhorn, S.K. Kunnath, C. Li and A. Madan, 6/3/96, (PB97-100234, A17, MF-A03).

NCEER-96-0011

Estimation of the Economic Impact of Multiple Lifeline Disruption: Memphis Light, Gas and Water
Division Case Study, by S.E. Chang, H.A. Seligson and R.T. Eguchi, 8/16/96, (PB97-133490, A11, MFA03).

NCEER-96-0012

Proceedings from the Sixth Japan-U.S. Workshop on Earthquake Resistant Design of Lifeline Facilities and
Countermeasures Against Soil Liquefaction, Edited by M. Hamada and T. ORourke, 9/11/96, (PB97133581, A99, MF-A06).

NCEER-96-0013

Chemical Hazards, Mitigation and Preparedness in Areas of High Seismic Risk: A Methodology for
Estimating the Risk of Post-Earthquake Hazardous Materials Release, by H.A. Seligson, R.T. Eguchi, K.J.
Tierney and K. Richmond, 11/7/96, (PB97-133565, MF-A02, A08).

NCEER-96-0014

Response of Steel Bridge Bearings to Reversed Cyclic Loading, by J.B. Mander, D-K. Kim, S.S. Chen and
G.J. Premus, 11/13/96, (PB97-140735, A12, MF-A03).

NCEER-96-0015

Highway Culvert Performance During Past Earthquakes, by T.L. Youd and C.J. Beckman, 11/25/96,
(PB97-133532, A06, MF-A01).

NCEER-97-0001

Evaluation, Prevention and Mitigation of Pounding Effects in Building Structures, by R.E. Valles and
A.M. Reinhorn, 2/20/97, (PB97-159552, A14, MF-A03).

NCEER-97-0002

Seismic Design Criteria for Bridges and Other Highway Structures, by C. Rojahn, R. Mayes, D.G.
Anderson, J. Clark, J.H. Hom, R.V. Nutt and M.J. ORourke, 4/30/97, (PB97-194658, A06, MF-A03).

422

NCEER-97-0003

Proceedings of the U.S.-Italian Workshop on Seismic Evaluation and Retrofit, Edited by D.P. Abrams and
G.M. Calvi, 3/19/97, (PB97-194666, A13, MF-A03).

NCEER-97-0004

"Investigation of Seismic Response of Buildings with Linear and Nonlinear Fluid Viscous Dampers," by
A.A. Seleemah and M.C. Constantinou, 5/21/97, (PB98-109002, A15, MF-A03).

NCEER-97-0005

"Proceedings of the Workshop on Earthquake Engineering Frontiers in Transportation Facilities," edited by


G.C. Lee and I.M. Friedland, 8/29/97, (PB98-128911, A25, MR-A04).

NCEER-97-0006

"Cumulative Seismic Damage of Reinforced Concrete Bridge Piers," by S.K. Kunnath, A. El-Bahy, A.
Taylor and W. Stone, 9/2/97, (PB98-108814, A11, MF-A03).

NCEER-97-0007

"Structural Details to Accommodate Seismic Movements of Highway Bridges and Retaining Walls," by R.A.
Imbsen, R.A. Schamber, E. Thorkildsen, A. Kartoum, B.T. Martin, T.N. Rosser and J.M. Kulicki, 9/3/97,
(PB98-108996, A09, MF-A02).

NCEER-97-0008

"A Method for Earthquake Motion-Damage Relationships with Application to Reinforced Concrete Frames,"
by A. Singhal and A.S. Kiremidjian, 9/10/97, (PB98-108988, A13, MF-A03).

NCEER-97-0009

"Seismic Analysis and Design of Bridge Abutments Considering Sliding and Rotation," by K. Fishman and
R. Richards, Jr., 9/15/97, (PB98-108897, A06, MF-A02).

NCEER-97-0010

"Proceedings of the FHWA/NCEER Workshop on the National Representation of Seismic Ground Motion
for New and Existing Highway Facilities," edited by I.M. Friedland, M.S. Power and R.L. Mayes, 9/22/97,
(PB98-128903, A21, MF-A04).

NCEER-97-0011

"Seismic Analysis for Design or Retrofit of Gravity Bridge Abutments," by K.L. Fishman, R. Richards, Jr.
and R.C. Divito, 10/2/97, (PB98-128937, A08, MF-A02).

NCEER-97-0012

"Evaluation of Simplified Methods of Analysis for Yielding Structures," by P. Tsopelas, M.C. Constantinou,
C.A. Kircher and A.S. Whittaker, 10/31/97, (PB98-128929, A10, MF-A03).

NCEER-97-0013

"Seismic Design of Bridge Columns Based on Control and Repairability of Damage," by C-T. Cheng and
J.B. Mander, 12/8/97, (PB98-144249, A11, MF-A03).

NCEER-97-0014

"Seismic Resistance of Bridge Piers Based on Damage Avoidance Design," by J.B. Mander and C-T. Cheng,
12/10/97, (PB98-144223, A09, MF-A02).

NCEER-97-0015

Seismic Response of Nominally Symmetric Systems with Strength Uncertainty, by S. Balopoulou and M.
Grigoriu, 12/23/97, (PB98-153422, A11, MF-A03).

NCEER-97-0016

Evaluation of Seismic Retrofit Methods for Reinforced Concrete Bridge Columns, by T.J. Wipf, F.W.
Klaiber and F.M. Russo, 12/28/97, (PB98-144215, A12, MF-A03).

NCEER-97-0017

Seismic Fragility of Existing Conventional Reinforced Concrete Highway Bridges, by C.L. Mullen and
A.S. Cakmak, 12/30/97, (PB98-153406, A08, MF-A02).

NCEER-97-0018

Loss Asssessment of Memphis Buildings, edited by D.P. Abrams and M. Shinozuka, 12/31/97, (PB98144231, A13, MF-A03).

NCEER-97-0019

Seismic Evaluation of Frames with Infill Walls Using Quasi-static Experiments, by K.M. Mosalam, R.N.
White and P. Gergely, 12/31/97, (PB98-153455, A07, MF-A02).

NCEER-97-0020

Seismic Evaluation of Frames with Infill Walls Using Pseudo-dynamic Experiments, by K.M. Mosalam,
R.N. White and P. Gergely, 12/31/97, (PB98-153430, A07, MF-A02).

NCEER-97-0021

Computational Strategies for Frames with Infill Walls: Discrete and Smeared Crack Analyses and Seismic
Fragility, by K.M. Mosalam, R.N. White and P. Gergely, 12/31/97, (PB98-153414, A10, MF-A02).

423

NCEER-97-0022

Proceedings of the NCEER Workshop on Evaluation of Liquefaction Resistance of Soils, edited by T.L.
Youd and I.M. Idriss, 12/31/97, (PB98-155617, A15, MF-A03).

MCEER-98-0001 Extraction of Nonlinear Hysteretic Properties of Seismically Isolated Bridges from Quick-Release Field
Tests, by Q. Chen, B.M. Douglas, E.M. Maragakis and I.G. Buckle, 5/26/98, (PB99-118838, A06, MFA01).
MCEER-98-0002 Methodologies for Evaluating the Importance of Highway Bridges, by A. Thomas, S. Eshenaur and J.
Kulicki, 5/29/98, (PB99-118846, A10, MF-A02).
MCEER-98-0003 Capacity Design of Bridge Piers and the Analysis of Overstrength, by J.B. Mander, A. Dutta and P. Goel,
6/1/98, (PB99-118853, A09, MF-A02).
MCEER-98-0004 Evaluation of Bridge Damage Data from the Loma Prieta and Northridge, California Earthquakes, by N.
Basoz and A. Kiremidjian, 6/2/98, (PB99-118861, A15, MF-A03).
MCEER-98-0005 Screening Guide for Rapid Assessment of Liquefaction Hazard at Highway Bridge Sites, by T. L. Youd,
6/16/98, (PB99-118879, A06, not available on microfiche).
MCEER-98-0006 Structural Steel and Steel/Concrete Interface Details for Bridges, by P. Ritchie, N. Kauhl and J. Kulicki,
7/13/98, (PB99-118945, A06, MF-A01).
MCEER-98-0007 Capacity Design and Fatigue Analysis of Confined Concrete Columns, by A. Dutta and J.B. Mander,
7/14/98, (PB99-118960, A14, MF-A03).
MCEER-98-0008 Proceedings of the Workshop on Performance Criteria for Telecommunication Services Under Earthquake
Conditions, edited by A.J. Schiff, 7/15/98, (PB99-118952, A08, MF-A02).
MCEER-98-0009 Fatigue Analysis of Unconfined Concrete Columns, by J.B. Mander, A. Dutta and J.H. Kim, 9/12/98,
(PB99-123655, A10, MF-A02).
MCEER-98-0010 Centrifuge Modeling of Cyclic Lateral Response of Pile-Cap Systems and Seat-Type Abutments in Dry
Sands, by A.D. Gadre and R. Dobry, 10/2/98, (PB99-123606, A13, MF-A03).
MCEER-98-0011 IDARC-BRIDGE: A Computational Platform for Seismic Damage Assessment of Bridge Structures, by
A.M. Reinhorn, V. Simeonov, G. Mylonakis and Y. Reichman, 10/2/98, (PB99-162919, A15, MF-A03).
MCEER-98-0012 Experimental Investigation of the Dynamic Response of Two Bridges Before and After Retrofitting with
Elastomeric Bearings, by D.A. Wendichansky, S.S. Chen and J.B. Mander, 10/2/98, (PB99-162927, A15,
MF-A03).
MCEER-98-0013 Design Procedures for Hinge Restrainers and Hinge Sear Width for Multiple-Frame Bridges, by R. Des
Roches and G.L. Fenves, 11/3/98, (PB99-140477, A13, MF-A03).
MCEER-98-0014

Response Modification Factors for Seismically Isolated Bridges, by M.C. Constantinou and J.K. Quarshie,
11/3/98, (PB99-140485, A14, MF-A03).

MCEER-98-0015

Proceedings of the U.S.-Italy Workshop on Seismic Protective Systems for Bridges, edited by I.M. Friedland
and M.C. Constantinou, 11/3/98, (PB2000-101711, A22, MF-A04).

MCEER-98-0016 Appropriate Seismic Reliability for Critical Equipment Systems: Recommendations Based on Regional
Analysis of Financial and Life Loss, by K. Porter, C. Scawthorn, C. Taylor and N. Blais, 11/10/98, (PB99157265, A08, MF-A02).
MCEER-98-0017 Proceedings of the U.S. Japan Joint Seminar on Civil Infrastructure Systems Research, edited by M.
Shinozuka and A. Rose, 11/12/98, (PB99-156713, A16, MF-A03).
MCEER-98-0018 Modeling of Pile Footings and Drilled Shafts for Seismic Design, by I. PoLam, M. Kapuskar and D.
Chaudhuri, 12/21/98, (PB99-157257, A09, MF-A02).

424

MCEER-99-0001 "Seismic Evaluation of a Masonry Infilled Reinforced Concrete Frame by Pseudodynamic Testing," by S.G.
Buonopane and R.N. White, 2/16/99, (PB99-162851, A09, MF-A02).
MCEER-99-0002 "Response History Analysis of Structures with Seismic Isolation and Energy Dissipation Systems:
Verification Examples for Program SAP2000," by J. Scheller and M.C. Constantinou, 2/22/99, (PB99162869, A08, MF-A02).
MCEER-99-0003 "Experimental Study on the Seismic Design and Retrofit of Bridge Columns Including Axial Load Effects,"
by A. Dutta, T. Kokorina and J.B. Mander, 2/22/99, (PB99-162877, A09, MF-A02).
MCEER-99-0004 "Experimental Study of Bridge Elastomeric and Other Isolation and Energy Dissipation Systems with
Emphasis on Uplift Prevention and High Velocity Near-source Seismic Excitation," by A. Kasalanati and M.
C. Constantinou, 2/26/99, (PB99-162885, A12, MF-A03).
MCEER-99-0005 "Truss Modeling of Reinforced Concrete Shear-flexure Behavior," by J.H. Kim and J.B. Mander, 3/8/99,
(PB99-163693, A12, MF-A03).
MCEER-99-0006 "Experimental Investigation and Computational Modeling of Seismic Response of a 1:4 Scale Model Steel
Structure with a Load Balancing Supplemental Damping System," by G. Pekcan, J.B. Mander and S.S. Chen,
4/2/99, (PB99-162893, A11, MF-A03).
MCEER-99-0007 "Effect of Vertical Ground Motions on the Structural Response of Highway Bridges," by M.R. Button, C.J.
Cronin and R.L. Mayes, 4/10/99, (PB2000-101411, A10, MF-A03).
MCEER-99-0008 "Seismic Reliability Assessment of Critical Facilities: A Handbook, Supporting Documentation, and Model
Code Provisions," by G.S. Johnson, R.E. Sheppard, M.D. Quilici, S.J. Eder and C.R. Scawthorn, 4/12/99,
(PB2000-101701, A18, MF-A04).
MCEER-99-0009 "Impact Assessment of Selected MCEER Highway Project Research on the Seismic Design of Highway
Structures," by C. Rojahn, R. Mayes, D.G. Anderson, J.H. Clark, D'Appolonia Engineering, S. Gloyd and
R.V. Nutt, 4/14/99, (PB99-162901, A10, MF-A02).
MCEER-99-0010 "Site Factors and Site Categories in Seismic Codes," by R. Dobry, R. Ramos and M.S. Power, 7/19/99,
(PB2000-101705, A08, MF-A02).
MCEER-99-0011 "Restrainer Design Procedures for Multi-Span Simply-Supported Bridges," by M.J. Randall, M. Saiidi, E.
Maragakis and T. Isakovic, 7/20/99, (PB2000-101702, A10, MF-A02).
MCEER-99-0012 "Property Modification Factors for Seismic Isolation Bearings," by M.C. Constantinou, P. Tsopelas, A.
Kasalanati and E. Wolff, 7/20/99, (PB2000-103387, A11, MF-A03).
MCEER-99-0013 "Critical Seismic Issues for Existing Steel Bridges," by P. Ritchie, N. Kauhl and J. Kulicki, 7/20/99,
(PB2000-101697, A09, MF-A02).
MCEER-99-0014 "Nonstructural Damage Database," by A. Kao, T.T. Soong and A. Vender, 7/24/99, (PB2000-101407, A06,
MF-A01).
MCEER-99-0015 "Guide to Remedial Measures for Liquefaction Mitigation at Existing Highway Bridge Sites," by H.G.
Cooke and J. K. Mitchell, 7/26/99, (PB2000-101703, A11, MF-A03).
MCEER-99-0016 "Proceedings of the MCEER Workshop on Ground Motion Methodologies for the Eastern United States,"
edited by N. Abrahamson and A. Becker, 8/11/99, (PB2000-103385, A07, MF-A02).
MCEER-99-0017 "Quindo, Colombia Earthquake of January 25, 1999: Reconnaissance Report," by A.P. Asfura and P.J.
Flores, 10/4/99, (PB2000-106893, A06, MF-A01).
MCEER-99-0018 "Hysteretic Models for Cyclic Behavior of Deteriorating Inelastic Structures," by M.V. Sivaselvan and A.M.
Reinhorn, 11/5/99, (PB2000-103386, A08, MF-A02).

425

MCEER-99-0019 "Proceedings of the 7th U.S.- Japan Workshop on Earthquake Resistant Design of Lifeline Facilities and
Countermeasures Against Soil Liquefaction," edited by T.D. O'Rourke, J.P. Bardet and M. Hamada,
11/19/99, (PB2000-103354, A99, MF-A06).
MCEER-99-0020 "Development of Measurement Capability for Micro-Vibration Evaluations with Application to Chip
Fabrication Facilities," by G.C. Lee, Z. Liang, J.W. Song, J.D. Shen and W.C. Liu, 12/1/99, (PB2000105993, A08, MF-A02).
MCEER-99-0021 "Design and Retrofit Methodology for Building Structures with Supplemental Energy Dissipating Systems,"
by G. Pekcan, J.B. Mander and S.S. Chen, 12/31/99, (PB2000-105994, A11, MF-A03).
MCEER-00-0001 "The Marmara, Turkey Earthquake of August 17, 1999: Reconnaissance Report," edited by C. Scawthorn;
with major contributions by M. Bruneau, R. Eguchi, T. Holzer, G. Johnson, J. Mander, J. Mitchell, W.
Mitchell, A. Papageorgiou, C. Scaethorn, and G. Webb, 3/23/00, (PB2000-106200, A11, MF-A03).
MCEER-00-0002 "Proceedings of the MCEER Workshop for Seismic Hazard Mitigation of Health Care Facilities," edited by
G.C. Lee, M. Ettouney, M. Grigoriu, J. Hauer and J. Nigg, 3/29/00, (PB2000-106892, A08, MF-A02).
MCEER-00-0003 "The Chi-Chi, Taiwan Earthquake of September 21, 1999: Reconnaissance Report," edited by G.C. Lee and
C.H. Loh, with major contributions by G.C. Lee, M. Bruneau, I.G. Buckle, S.E. Chang, P.J. Flores, T.D.
O'Rourke, M. Shinozuka, T.T. Soong, C-H. Loh, K-C. Chang, Z-J. Chen, J-S. Hwang, M-L. Lin, G-Y. Liu,
K-C. Tsai, G.C. Yao and C-L. Yen, 4/30/00, (PB2001-100980, A10, MF-A02).
MCEER-00-0004 "Seismic Retrofit of End-Sway Frames of Steel Deck-Truss Bridges with a Supplemental Tendon System:
Experimental and Analytical Investigation," by G. Pekcan, J.B. Mander and S.S. Chen, 7/1/00, (PB2001100982, A10, MF-A02).
MCEER-00-0005 "Sliding Fragility of Unrestrained Equipment in Critical Facilities," by W.H. Chong and T.T. Soong, 7/5/00,
(PB2001-100983, A08, MF-A02).
MCEER-00-0006 "Seismic Response of Reinforced Concrete Bridge Pier Walls in the Weak Direction," by N. Abo-Shadi, M.
Saiidi and D. Sanders, 7/17/00, (PB2001-100981, A17, MF-A03).
MCEER-00-0007 "Low-Cycle Fatigue Behavior of Longitudinal Reinforcement in Reinforced Concrete Bridge Columns," by
J. Brown and S.K. Kunnath, 7/23/00, (PB2001-104392, A08, MF-A02).
MCEER-00-0008 "Soil Structure Interaction of Bridges for Seismic Analysis," I. PoLam and H. Law, 9/25/00, (PB2001105397, A08, MF-A02).
MCEER-00-0009 "Proceedings of the First MCEER Workshop on Mitigation of Earthquake Disaster by Advanced
Technologies (MEDAT-1), edited by M. Shinozuka, D.J. Inman and T.D. O'Rourke, 11/10/00, (PB2001105399, A14, MF-A03).
MCEER-00-0010 "Development and Evaluation of Simplified Procedures for Analysis and Design of Buildings with Passive
Energy Dissipation Systems, Revision 01," by O.M. Ramirez, M.C. Constantinou, C.A. Kircher, A.S.
Whittaker, M.W. Johnson, J.D. Gomez and C. Chrysostomou, 11/16/01, (PB2001-105523, A23, MF-A04).
MCEER-00-0011 "Dynamic Soil-Foundation-Structure Interaction Analyses of Large Caissons," by C-Y. Chang, C-M. Mok,
Z-L. Wang, R. Settgast, F. Waggoner, M.A. Ketchum, H.M. Gonnermann and C-C. Chin, 12/30/00,
(PB2001-104373, A07, MF-A02).
MCEER-00-0012 "Experimental Evaluation of Seismic Performance of Bridge Restrainers," by A.G. Vlassis, E.M. Maragakis
and M. Saiid Saiidi, 12/30/00, (PB2001-104354, A09, MF-A02).
MCEER-00-0013 "Effect of Spatial Variation of Ground Motion on Highway Structures," by M. Shinozuka, V. Saxena and G.
Deodatis, 12/31/00, (PB2001-108755, A13, MF-A03).
MCEER-00-0014 "A Risk-Based Methodology for Assessing the Seismic Performance of Highway Systems," by S.D. Werner,
C.E. Taylor, J.E. Moore, II, J.S. Walton and S. Cho, 12/31/00, (PB2001-108756, A14, MF-A03).

426

MCEER-01-0001 Experimental Investigation of P-Delta Effects to Collapse During Earthquakes, by D. Vian and M.
Bruneau, 6/25/01, (PB2002-100534, A17, MF-A03).
MCEER-01-0002 Proceedings of the Second MCEER Workshop on Mitigation of Earthquake Disaster by Advanced
Technologies (MEDAT-2), edited by M. Bruneau and D.J. Inman, 7/23/01, (PB2002-100434, A16, MFA03).
MCEER-01-0003 Sensitivity Analysis of Dynamic Systems Subjected to Seismic Loads, by C. Roth and M. Grigoriu,
9/18/01, (PB2003-100884, A12, MF-A03).
MCEER-01-0004 Overcoming Obstacles to Implementing Earthquake Hazard Mitigation Policies: Stage 1 Report, by D.J.
Alesch and W.J. Petak, 12/17/01, (PB2002-107949, A07, MF-A02).
MCEER-01-0005 Updating Real-Time Earthquake Loss Estimates: Methods, Problems and Insights, by C.E. Taylor, S.E.
Chang and R.T. Eguchi, 12/17/01, (PB2002-107948, A05, MF-A01).
MCEER-01-0006 Experimental Investigation and Retrofit of Steel Pile Foundations and Pile Bents Under Cyclic Lateral
Loadings, by A. Shama, J. Mander, B. Blabac and S. Chen, 12/31/01, (PB2002-107950, A13, MF-A03).
MCEER-02-0001 Assessment of Performance of Bolu Viaduct in the 1999 Duzce Earthquake in Turkey by P.C. Roussis,
M.C. Constantinou, M. Erdik, E. Durukal and M. Dicleli, 5/8/02, (PB2003-100883, A08, MF-A02).
MCEER-02-0002 Seismic Behavior of Rail Counterweight Systems of Elevators in Buildings, by M.P. Singh, Rildova and
L.E. Suarez, 5/27/02. (PB2003-100882, A11, MF-A03).
MCEER-02-0003 Development of Analysis and Design Procedures for Spread Footings, by G. Mylonakis, G. Gazetas, S.
Nikolaou and A. Chauncey, 10/02/02, (PB2004-101636, A13, MF-A03, CD-A13).
MCEER-02-0004 Bare-Earth Algorithms for Use with SAR and LIDAR Digital Elevation Models, by C.K. Huyck, R.T.
Eguchi and B. Houshmand, 10/16/02, (PB2004-101637, A07, CD-A07).
MCEER-02-0005 Review of Energy Dissipation of Compression Members in Concentrically Braced Frames, by K.Lee and
M. Bruneau, 10/18/02, (PB2004-101638, A10, CD-A10).
MCEER-03-0001 Experimental Investigation of Light-Gauge Steel Plate Shear Walls for the Seismic Retrofit of Buildings
by J. Berman and M. Bruneau, 5/2/03, (PB2004-101622, A10, MF-A03, CD-A10).
MCEER-03-0002 Statistical Analysis of Fragility Curves, by M. Shinozuka, M.Q. Feng, H. Kim, T. Uzawa and T. Ueda,
6/16/03, (PB2004-101849, A09, CD-A09).
MCEER-03-0003 Proceedings of the Eighth U.S.-Japan Workshop on Earthquake Resistant Design f Lifeline Facilities and
Countermeasures Against Liquefaction, edited by M. Hamada, J.P. Bardet and T.D. ORourke, 6/30/03,
(PB2004-104386, A99, CD-A99).
MCEER-03-0004 Proceedings of the PRC-US Workshop on Seismic Analysis and Design of Special Bridges, edited by L.C.
Fan and G.C. Lee, 7/15/03, (PB2004-104387, A14, CD-A14).
MCEER-03-0005 Urban Disaster Recovery: A Framework and Simulation Model, by S.B. Miles and S.E. Chang, 7/25/03,
(PB2004-104388, A07, CD-A07).
MCEER-03-0006 Behavior of Underground Piping Joints Due to Static and Dynamic Loading, by R.D. Meis, M. Maragakis
and R. Siddharthan, 11/17/03, (PB2005-102194, A13, MF-A03, CD-A00).
MCEER-04-0001 Experimental Study of Seismic Isolation Systems with Emphasis on Secondary System Response and
Verification of Accuracy of Dynamic Response History Analysis Methods, by E. Wolff and M.
Constantinou, 1/16/04 (PB2005-102195, A99, MF-E08, CD-A00).
MCEER-04-0002 Tension, Compression and Cyclic Testing of Engineered Cementitious Composite Materials, by K. Kesner
and S.L. Billington, 3/1/04, (PB2005-102196, A08, CD-A08).

427

MCEER-04-0003 Cyclic Testing of Braces Laterally Restrained by Steel Studs to Enhance Performance During Earthquakes,
by O.C. Celik, J.W. Berman and M. Bruneau, 3/16/04, (PB2005-102197, A13, MF-A03, CD-A00).
MCEER-04-0004 Methodologies for Post Earthquake Building Damage Detection Using SAR and Optical Remote Sensing:
Application to the August 17, 1999 Marmara, Turkey Earthquake, by C.K. Huyck, B.J. Adams, S. Cho,
R.T. Eguchi, B. Mansouri and B. Houshmand, 6/15/04, (PB2005-104888, A10, CD-A00).
MCEER-04-0005 Nonlinear Structural Analysis Towards Collapse Simulation: A Dynamical Systems Approach, by M.V.
Sivaselvan and A.M. Reinhorn, 6/16/04, (PB2005-104889, A11, MF-A03, CD-A00).
MCEER-04-0006 Proceedings of the Second PRC-US Workshop on Seismic Analysis and Design of Special Bridges, edited
by G.C. Lee and L.C. Fan, 6/25/04, (PB2005-104890, A16, CD-A00).
MCEER-04-0007 Seismic Vulnerability Evaluation of Axially Loaded Steel Built-up Laced Members, by K. Lee and M.
Bruneau, 6/30/04, (PB2005-104891, A16, CD-A00).
MCEER-04-0008 Evaluation of Accuracy of Simplified Methods of Analysis and Design of Buildings with Damping Systems
for Near-Fault and for Soft-Soil Seismic Motions, by E.A. Pavlou and M.C. Constantinou, 8/16/04,
(PB2005-104892, A08, MF-A02, CD-A00).
MCEER-04-0009 Assessment of Geotechnical Issues in Acute Care Facilities in California, by M. Lew, T.D. ORourke, R.
Dobry and M. Koch, 9/15/04, (PB2005-104893, A08, CD-A00).
MCEER-04-0010 Scissor-Jack-Damper Energy Dissipation System, by A.N. Sigaher-Boyle and M.C. Constantinou, 12/1/04
(PB2005-108221).
MCEER-04-0011 Seismic Retrofit of Bridge Steel Truss Piers Using a Controlled Rocking Approach, by M. Pollino and M.
Bruneau, 12/20/04 (PB2006-105795).
MCEER-05-0001 Experimental and Analytical Studies of Structures Seismically Isolated with an Uplift-Restraint Isolation
System, by P.C. Roussis and M.C. Constantinou, 1/10/05 (PB2005-108222).
MCEER-05-0002 A Versatile Experimentation Model for Study of Structures Near Collapse Applied to Seismic Evaluation of
Irregular Structures, by D. Kusumastuti, A.M. Reinhorn and A. Rutenberg, 3/31/05 (PB2006-101523).
MCEER-05-0003 Proceedings of the Third PRC-US Workshop on Seismic Analysis and Design of Special Bridges, edited
by L.C. Fan and G.C. Lee, 4/20/05, (PB2006-105796).
MCEER-05-0004 Approaches for the Seismic Retrofit of Braced Steel Bridge Piers and Proof-of-Concept Testing of an
Eccentrically Braced Frame with Tubular Link, by J.W. Berman and M. Bruneau, 4/21/05 (PB2006101524).
MCEER-05-0005 Simulation of Strong Ground Motions for Seismic Fragility Evaluation of Nonstructural Components in
Hospitals, by A. Wanitkorkul and A. Filiatrault, 5/26/05 (PB2006-500027).
MCEER-05-0006 Seismic Safety in California Hospitals: Assessing an Attempt to Accelerate the Replacement or Seismic
Retrofit of Older Hospital Facilities, by D.J. Alesch, L.A. Arendt and W.J. Petak, 6/6/05 (PB2006-105794).
MCEER-05-0007 Development of Seismic Strengthening and Retrofit Strategies for Critical Facilities Using Engineered
Cementitious Composite Materials, by K. Kesner and S.L. Billington, 8/29/05 (PB2006-111701).
MCEER-05-0008 Experimental and Analytical Studies of Base Isolation Systems for Seismic Protection of Power
Transformers, by N. Murota, M.Q. Feng and G-Y. Liu, 9/30/05 (PB2006-111702).
MCEER-05-0009 3D-BASIS-ME-MB: Computer Program for Nonlinear Dynamic Analysis of Seismically Isolated
Structures, by P.C. Tsopelas, P.C. Roussis, M.C. Constantinou, R. Buchanan and A.M. Reinhorn, 10/3/05
(PB2006-111703).
MCEER-05-0010 Steel Plate Shear Walls for Seismic Design and Retrofit of Building Structures, by D. Vian and M.
Bruneau, 12/15/05 (PB2006-111704).

428

MCEER-05-0011 The Performance-Based Design Paradigm, by M.J. Astrella and A. Whittaker, 12/15/05 (PB2006-111705).
MCEER-06-0001 Seismic Fragility of Suspended Ceiling Systems, H. Badillo-Almaraz, A.S. Whittaker, A.M. Reinhorn and
G.P. Cimellaro, 2/4/06 (PB2006-111706).
MCEER-06-0002 Multi-Dimensional Fragility of Structures, by G.P. Cimellaro, A.M. Reinhorn and M. Bruneau, 3/1/06
(PB2007-106974, A09, MF-A02, CD A00).
MCEER-06-0003 Built-Up Shear Links as Energy Dissipators for Seismic Protection of Bridges, by P. Dusicka, A.M. Itani
and I.G. Buckle, 3/15/06 (PB2006-111708).
MCEER-06-0004 Analytical Investigation of the Structural Fuse Concept, by R.E. Vargas and M. Bruneau, 3/16/06
(PB2006-111709).
MCEER-06-0005 Experimental Investigation of the Structural Fuse Concept, by R.E. Vargas and M. Bruneau, 3/17/06
(PB2006-111710).
MCEER-06-0006 Further Development of Tubular Eccentrically Braced Frame Links for the Seismic Retrofit of Braced Steel
Truss Bridge Piers, by J.W. Berman and M. Bruneau, 3/27/06 (PB2007-105147).
MCEER-06-0007 REDARS Validation Report, by S. Cho, C.K. Huyck, S. Ghosh and R.T. Eguchi, 8/8/06 (PB2007-106983).
MCEER-06-0008 Review of Current NDE Technologies for Post-Earthquake Assessment of Retrofitted Bridge Columns, by
J.W. Song, Z. Liang and G.C. Lee, 8/21/06 (PB2007-106984).
MCEER-06-0009 Liquefaction Remediation in Silty Soils Using Dynamic Compaction and Stone Columns, by S.
Thevanayagam, G.R. Martin, R. Nashed, T. Shenthan, T. Kanagalingam and N. Ecemis, 8/28/06 (PB2007106985).
MCEER-06-0010 Conceptual Design and Experimental Investigation of Polymer Matrix Composite Infill Panels for Seismic
Retrofitting, by W. Jung, M. Chiewanichakorn and A.J. Aref, 9/21/06 (PB2007-106986).
MCEER-06-0011 A Study of the Coupled Horizontal-Vertical Behavior of Elastomeric and Lead-Rubber Seismic Isolation
Bearings, by G.P. Warn and A.S. Whittaker, 9/22/06 (PB2007-108679).
MCEER-06-0012 Proceedings of the Fourth PRC-US Workshop on Seismic Analysis and Design of Special Bridges:
Advancing Bridge Technologies in Research, Design, Construction and Preservation, Edited by L.C. Fan,
G.C. Lee and L. Ziang, 10/12/06 (PB2007-109042).
MCEER-06-0013 Cyclic Response and Low Cycle Fatigue Characteristics of Plate Steels, by P. Dusicka, A.M. Itani and I.G.
Buckle, 11/1/06 06 (PB2007-106987).
MCEER-06-0014 Proceedings of the Second US-Taiwan Bridge Engineering Workshop, edited by W.P. Yen, J. Shen, J-Y.
Chen and M. Wang, 11/15/06 (PB2008-500041).
MCEER-06-0015 User Manual and Technical Documentation for the REDARSTM Import Wizard, by S. Cho, S. Ghosh, C.K.
Huyck and S.D. Werner, 11/30/06 (PB2007-114766).
MCEER-06-0016 Hazard Mitigation Strategy and Monitoring Technologies for Urban and Infrastructure Public Buildings:
Proceedings of the China-US Workshops, edited by X.Y. Zhou, A.L. Zhang, G.C. Lee and M. Tong,
12/12/06 (PB2008-500018).
MCEER-07-0001 Static and Kinetic Coefficients of Friction for Rigid Blocks, by C. Kafali, S. Fathali, M. Grigoriu and A.S.
Whittaker, 3/20/07 (PB2007-114767).
MCEER-07-0002 Hazard Mitigation Investment Decision Making: Organizational Response to Legislative Mandate, by L.A.
Arendt, D.J. Alesch and W.J. Petak, 4/9/07 (PB2007-114768).
MCEER-07-0003 Seismic Behavior of Bidirectional-Resistant Ductile End Diaphragms with Unbonded Braces in Straight or
Skewed Steel Bridges, by O. Celik and M. Bruneau, 4/11/07 (PB2008-105141).

429

MCEER-07-0004 Modeling Pile Behavior in Large Pile Groups Under Lateral Loading, by A.M. Dodds and G.R. Martin,
4/16/07(PB2008-105142).
MCEER-07-0005 Experimental Investigation of Blast Performance of Seismically Resistant Concrete-Filled Steel Tube
Bridge Piers, by S. Fujikura, M. Bruneau and D. Lopez-Garcia, 4/20/07 (PB2008-105143).
MCEER-07-0006 Seismic Analysis of Conventional and Isolated Liquefied Natural Gas Tanks Using Mechanical Analogs,
by I.P. Christovasilis and A.S. Whittaker, 5/1/07, not available.
MCEER-07-0007 Experimental Seismic Performance Evaluation of Isolation/Restraint Systems for Mechanical Equipment
Part 1: Heavy Equipment Study, by S. Fathali and A. Filiatrault, 6/6/07 (PB2008-105144).
MCEER-07-0008 Seismic Vulnerability of Timber Bridges and Timber Substructures, by A.A. Sharma, J.B. Mander, I.M.
Friedland and D.R. Allicock, 6/7/07 (PB2008-105145).
MCEER-07-0009 Experimental and Analytical Study of the XY-Friction Pendulum (XY-FP) Bearing for Bridge
Applications, by C.C. Marin-Artieda, A.S. Whittaker and M.C. Constantinou, 6/7/07 (PB2008-105191).
MCEER-07-0010 Proceedings of the PRC-US Earthquake Engineering Forum for Young Researchers, Edited by G.C. Lee
and X.Z. Qi, 6/8/07 (PB2008-500058).
MCEER-07-0011 Design Recommendations for Perforated Steel Plate Shear Walls, by R. Purba and M. Bruneau, 6/18/07,
(PB2008-105192).
MCEER-07-0012 Performance of Seismic Isolation Hardware Under Service and Seismic Loading, by M.C. Constantinou,
A.S. Whittaker, Y. Kalpakidis, D.M. Fenz and G.P. Warn, 8/27/07, (PB2008-105193).
MCEER-07-0013 Experimental Evaluation of the Seismic Performance of Hospital Piping Subassemblies, by E.R. Goodwin,
E. Maragakis and A.M. Itani, 9/4/07, (PB2008-105194).
MCEER-07-0014 A Simulation Model of Urban Disaster Recovery and Resilience: Implementation for the 1994 Northridge
Earthquake, by S. Miles and S.E. Chang, 9/7/07, (PB2008-106426).
MCEER-07-0015 Statistical and Mechanistic Fragility Analysis of Concrete Bridges, by M. Shinozuka, S. Banerjee and S-H.
Kim, 9/10/07, (PB2008-106427).
MCEER-07-0016 Three-Dimensional Modeling of Inelastic Buckling in Frame Structures, by M. Schachter and AM.
Reinhorn, 9/13/07, (PB2008-108125).
MCEER-07-0017 Modeling of Seismic Wave Scattering on Pile Groups and Caissons, by I. Po Lam, H. Law and C.T. Yang,
9/17/07 (PB2008-108150).
MCEER-07-0018 Bridge Foundations: Modeling Large Pile Groups and Caissons for Seismic Design, by I. Po Lam, H. Law
and G.R. Martin (Coordinating Author), 12/1/07 (PB2008-111190).
MCEER-07-0019 Principles and Performance of Roller Seismic Isolation Bearings for Highway Bridges, by G.C. Lee, Y.C.
Ou, Z. Liang, T.C. Niu and J. Song, 12/10/07 (PB2009-110466).
MCEER-07-0020 Centrifuge Modeling of Permeability and Pinning Reinforcement Effects on Pile Response to Lateral
Spreading, by L.L Gonzalez-Lagos, T. Abdoun and R. Dobry, 12/10/07 (PB2008-111191).
MCEER-07-0021 Damage to the Highway System from the Pisco, Per Earthquake of August 15, 2007, by J.S. OConnor,
L. Mesa and M. Nykamp, 12/10/07, (PB2008-108126).
MCEER-07-0022 Experimental Seismic Performance Evaluation of Isolation/Restraint Systems for Mechanical Equipment
Part 2: Light Equipment Study, by S. Fathali and A. Filiatrault, 12/13/07 (PB2008-111192).
MCEER-07-0023 Fragility Considerations in Highway Bridge Design, by M. Shinozuka, S. Banerjee and S.H. Kim, 12/14/07
(PB2008-111193).

430

MCEER-07-0024 Performance Estimates for Seismically Isolated Bridges, by G.P. Warn and A.S. Whittaker, 12/30/07
(PB2008-112230).
MCEER-08-0001 Seismic Performance of Steel Girder Bridge Superstructures with Conventional Cross Frames, by L.P.
Carden, A.M. Itani and I.G. Buckle, 1/7/08, (PB2008-112231).
MCEER-08-0002 Seismic Performance of Steel Girder Bridge Superstructures with Ductile End Cross Frames with Seismic
Isolators, by L.P. Carden, A.M. Itani and I.G. Buckle, 1/7/08 (PB2008-112232).
MCEER-08-0003 Analytical and Experimental Investigation of a Controlled Rocking Approach for Seismic Protection of
Bridge Steel Truss Piers, by M. Pollino and M. Bruneau, 1/21/08 (PB2008-112233).
MCEER-08-0004 Linking Lifeline Infrastructure Performance and Community Disaster Resilience: Models and MultiStakeholder Processes, by S.E. Chang, C. Pasion, K. Tatebe and R. Ahmad, 3/3/08 (PB2008-112234).
MCEER-08-0005 Modal Analysis of Generally Damped Linear Structures Subjected to Seismic Excitations, by J. Song, Y-L.
Chu, Z. Liang and G.C. Lee, 3/4/08 (PB2009-102311).
MCEER-08-0006 System Performance Under Multi-Hazard Environments, by C. Kafali and M. Grigoriu, 3/4/08 (PB2008112235).
MCEER-08-0007 Mechanical Behavior of Multi-Spherical Sliding Bearings, by D.M. Fenz and M.C. Constantinou, 3/6/08
(PB2008-112236).
MCEER-08-0008 Post-Earthquake Restoration of the Los Angeles Water Supply System, by T.H.P. Tabucchi and R.A.
Davidson, 3/7/08 (PB2008-112237).
MCEER-08-0009 Fragility Analysis of Water Supply Systems, by A. Jacobson and M. Grigoriu, 3/10/08 (PB2009-105545).
MCEER-08-0010 Experimental Investigation of Full-Scale Two-Story Steel Plate Shear Walls with Reduced Beam Section
Connections, by B. Qu, M. Bruneau, C-H. Lin and K-C. Tsai, 3/17/08 (PB2009-106368).
MCEER-08-0011 Seismic Evaluation and Rehabilitation of Critical Components of Electrical Power Systems, S. Ersoy, B.
Feizi, A. Ashrafi and M. Ala Saadeghvaziri, 3/17/08 (PB2009-105546).
MCEER-08-0012 Seismic Behavior and Design of Boundary Frame Members of Steel Plate Shear Walls, by B. Qu and M.
Bruneau, 4/26/08 . (PB2009-106744).
MCEER-08-0013 Development and Appraisal of a Numerical Cyclic Loading Protocol for Quantifying Building System
Performance, by A. Filiatrault, A. Wanitkorkul and M. Constantinou, 4/27/08 (PB2009-107906).
MCEER-08-0014 Structural and Nonstructural Earthquake Design: The Challenge of Integrating Specialty Areas in Designing
Complex, Critical Facilities, by W.J. Petak and D.J. Alesch, 4/30/08 (PB2009-107907).
MCEER-08-0015 Seismic Performance Evaluation of Water Systems, by Y. Wang and T.D. ORourke, 5/5/08 (PB2009107908).
MCEER-08-0016 Seismic Response Modeling of Water Supply Systems, by P. Shi and T.D. ORourke, 5/5/08 (PB2009107910).
MCEER-08-0017 Numerical and Experimental Studies of Self-Centering Post-Tensioned Steel Frames, by D. Wang and A.
Filiatrault, 5/12/08 (PB2009-110479).
MCEER-08-0018 Development, Implementation and Verification of Dynamic Analysis Models for Multi-Spherical Sliding
Bearings, by D.M. Fenz and M.C. Constantinou, 8/15/08 (PB2009-107911).
MCEER-08-0019 Performance Assessment of Conventional and Base Isolated Nuclear Power Plants for Earthquake Blast
Loadings, by Y.N. Huang, A.S. Whittaker and N. Luco, 10/28/08 (PB2009-107912).

431

MCEER-08-0020 Remote Sensing for Resilient Multi-Hazard Disaster Response Volume I: Introduction to Damage
Assessment Methodologies, by B.J. Adams and R.T. Eguchi, 11/17/08 (PB2010-102695).
MCEER-08-0021 Remote Sensing for Resilient Multi-Hazard Disaster Response Volume II: Counting the Number of
Collapsed Buildings Using an Object-Oriented Analysis: Case Study of the 2003 Bam Earthquake, by L.
Gusella, C.K. Huyck and B.J. Adams, 11/17/08 (PB2010-100925).
MCEER-08-0022 Remote Sensing for Resilient Multi-Hazard Disaster Response Volume III: Multi-Sensor Image Fusion
Techniques for Robust Neighborhood-Scale Urban Damage Assessment, by B.J. Adams and A. McMillan,
11/17/08 (PB2010-100926).
MCEER-08-0023 Remote Sensing for Resilient Multi-Hazard Disaster Response Volume IV: A Study of Multi-Temporal
and Multi-Resolution SAR Imagery for Post-Katrina Flood Monitoring in New Orleans, by A. McMillan,
J.G. Morley, B.J. Adams and S. Chesworth, 11/17/08 (PB2010-100927).
MCEER-08-0024 Remote Sensing for Resilient Multi-Hazard Disaster Response Volume V: Integration of Remote Sensing
Imagery and VIEWSTM Field Data for Post-Hurricane Charley Building Damage Assessment, by J.A.
Womble, K. Mehta and B.J. Adams, 11/17/08 (PB2009-115532).
MCEER-08-0025 Building Inventory Compilation for Disaster Management: Application of Remote Sensing and Statistical
Modeling, by P. Sarabandi, A.S. Kiremidjian, R.T. Eguchi and B. J. Adams, 11/20/08 (PB2009-110484).
MCEER-08-0026 New Experimental Capabilities and Loading Protocols for Seismic Qualification and Fragility Assessment
of Nonstructural Systems, by R. Retamales, G. Mosqueda, A. Filiatrault and A. Reinhorn, 11/24/08
(PB2009-110485).
MCEER-08-0027 Effects of Heating and Load History on the Behavior of Lead-Rubber Bearings, by I.V. Kalpakidis and
M.C. Constantinou, 12/1/08 (PB2009-115533).
MCEER-08-0028 Experimental and Analytical Investigation of Blast Performance of Seismically Resistant Bridge Piers, by
S.Fujikura and M. Bruneau, 12/8/08 (PB2009-115534).
MCEER-08-0029 Evolutionary Methodology for Aseismic Decision Support, by Y. Hu and G. Dargush, 12/15/08.
MCEER-08-0030 Development of a Steel Plate Shear Wall Bridge Pier System Conceived from a Multi-Hazard Perspective,
by D. Keller and M. Bruneau, 12/19/08 (PB2010-102696).
MCEER-09-0001 Modal Analysis of Arbitrarily Damped Three-Dimensional Linear Structures Subjected to Seismic
Excitations, by Y.L. Chu, J. Song and G.C. Lee, 1/31/09 (PB2010-100922).
MCEER-09-0002 Air-Blast Effects on Structural Shapes, by G. Ballantyne, A.S. Whittaker, A.J. Aref and G.F. Dargush,
2/2/09 (PB2010-102697).
MCEER-09-0003 Water Supply Performance During Earthquakes and Extreme Events, by A.L. Bonneau and T.D.
ORourke, 2/16/09 (PB2010-100923).
MCEER-09-0004 Generalized Linear (Mixed) Models of Post-Earthquake Ignitions, by R.A. Davidson, 7/20/09 (PB2010102698).
MCEER-09-0005 Seismic Testing of a Full-Scale Two-Story Light-Frame Wood Building: NEESWood Benchmark Test, by
I.P. Christovasilis, A. Filiatrault and A. Wanitkorkul, 7/22/09 (PB2012-102401).
MCEER-09-0006 IDARC2D Version 7.0: A Program for the Inelastic Damage Analysis of Structures, by A.M. Reinhorn, H.
Roh, M. Sivaselvan, S.K. Kunnath, R.E. Valles, A. Madan, C. Li, R. Lobo and Y.J. Park, 7/28/09 (PB2010103199).
MCEER-09-0007 Enhancements to Hospital Resiliency: Improving Emergency Planning for and Response to Hurricanes, by
D.B. Hess and L.A. Arendt, 7/30/09 (PB2010-100924).

432

MCEER-09-0008 Assessment of Base-Isolated Nuclear Structures for Design and Beyond-Design Basis Earthquake Shaking,
by Y.N. Huang, A.S. Whittaker, R.P. Kennedy and R.L. Mayes, 8/20/09 (PB2010-102699).
MCEER-09-0009 Quantification of Disaster Resilience of Health Care Facilities, by G.P. Cimellaro, C. Fumo, A.M Reinhorn
and M. Bruneau, 9/14/09 (PB2010-105384).
MCEER-09-0010 Performance-Based Assessment and Design of Squat Reinforced Concrete Shear Walls, by C.K. Gulec and
A.S. Whittaker, 9/15/09 (PB2010-102700).
MCEER-09-0011 Proceedings of the Fourth US-Taiwan Bridge Engineering Workshop, edited by W.P. Yen, J.J. Shen, T.M.
Lee and R.B. Zheng, 10/27/09 (PB2010-500009).
MCEER-09-0012 Proceedings of the Special International Workshop on Seismic Connection Details for Segmental Bridge
Construction, edited by W. Phillip Yen and George C. Lee, 12/21/09 (PB2012-102402).
MCEER-10-0001 Direct Displacement Procedure for Performance-Based Seismic Design of Multistory Woodframe
Structures, by W. Pang and D. Rosowsky, 4/26/10 (PB2012-102403).
MCEER-10-0002 Simplified Direct Displacement Design of Six-Story NEESWood Capstone Building and Pre-Test Seismic
Performance Assessment, by W. Pang, D. Rosowsky, J. van de Lindt and S. Pei, 5/28/10 (PB2012-102404).
MCEER-10-0003 Integration of Seismic Protection Systems in Performance-Based Seismic Design of Woodframed
Structures, by J.K. Shinde and M.D. Symans, 6/18/10 (PB2012-102405).
MCEER-10-0004 Modeling and Seismic Evaluation of Nonstructural Components: Testing Frame for Experimental
Evaluation of Suspended Ceiling Systems, by A.M. Reinhorn, K.P. Ryu and G. Maddaloni, 6/30/10
(PB2012-102406).
MCEER-10-0005 Analytical Development and Experimental Validation of a Structural-Fuse Bridge Pier Concept, by S. ElBahey and M. Bruneau, 10/1/10 (PB2012-102407).
MCEER-10-0006 A Framework for Defining and Measuring Resilience at the Community Scale: The PEOPLES Resilience
Framework, by C.S. Renschler, A.E. Frazier, L.A. Arendt, G.P. Cimellaro, A.M. Reinhorn and M. Bruneau,
10/8/10 (PB2012-102408).
MCEER-10-0007 Impact of Horizontal Boundary Elements Design on Seismic Behavior of Steel Plate Shear Walls, by R.
Purba and M. Bruneau, 11/14/10 (PB2012-102409).
MCEER-10-0008 Seismic Testing of a Full-Scale Mid-Rise Building: The NEESWood Capstone Test, by S. Pei, J.W. van de
Lindt, S.E. Pryor, H. Shimizu, H. Isoda and D.R. Rammer, 12/1/10 (PB2012-102410).
MCEER-10-0009 Modeling the Effects of Detonations of High Explosives to Inform Blast-Resistant Design, by P. Sherkar,
A.S. Whittaker and A.J. Aref, 12/1/10 (PB2012-102411).
MCEER-10-0010 LAquila Earthquake of April 6, 2009 in Italy: Rebuilding a Resilient City to Withstand Multiple Hazards,
by G.P. Cimellaro, I.P. Christovasilis, A.M. Reinhorn, A. De Stefano and T. Kirova, 12/29/10.
MCEER-11-0001 Numerical and Experimental Investigation of the Seismic Response of Light-Frame Wood Structures, by
I.P. Christovasilis and A. Filiatrault, 8/8/11 (PB2012-102412).
MCEER-11-0002 Seismic Design and Analysis of a Precast Segmental Concrete Bridge Model, by M. Anagnostopoulou, A.
Filiatrault and A. Aref, 9/15/11.
MCEER-11-0003 Proceedings of the Workshop on Improving Earthquake Response of Substation Equipment, Edited by
A.M. Reinhorn, 9/19/11 (PB2012-102413).
MCEER-11-0004 LRFD-Based Analysis and Design Procedures for Bridge Bearings and Seismic Isolators, by M.C.
Constantinou, I. Kalpakidis, A. Filiatrault and R.A. Ecker Lay, 9/26/11.

433

MCEER-11-0005 Experimental Seismic Evaluation, Model Parameterization, and Effects of Cold-Formed Steel-Framed
Gypsum Partition Walls on the Seismic Performance of an Essential Facility, by R. Davies, R. Retamales,
G. Mosqueda and A. Filiatrault, 10/12/11.
MCEER-11-0006 Modeling and Seismic Performance Evaluation of High Voltage Transformers and Bushings, by A.M.
Reinhorn, K. Oikonomou, H. Roh, A. Schiff and L. Kempner, Jr., 10/3/11.
MCEER-11-0007 Extreme Load Combinations: A Survey of State Bridge Engineers, by G.C. Lee, Z. Liang, J.J. Shen and
J.S. OConnor, 10/14/11.
MCEER-12-0001 Simplified Analysis Procedures in Support of Performance Based Seismic Design, by Y.N. Huang and
A.S. Whittaker.
MCEER-12-0002 Seismic Protection of Electrical Transformer Bushing Systems by Stiffening Techniques, by M. Koliou, A.
Filiatrault, A.M. Reinhorn and N. Oliveto, 6/1/12.
MCEER-12-0003 Post-Earthquake Bridge Inspection Guidelines, by J.S. OConnor and S. Alampalli, 6/8/12.
MCEER-12-0004 Integrated Design Methodology for Isolated Floor Systems in Single-Degree-of-Freedom Structural Fuse
Systems, by S. Cui, M. Bruneau and M.C. Constantinou, 6/13/12.
MCEER-12-0005 Characterizing the Rotational Components of Earthquake Ground Motion, by D. Basu, A.S. Whittaker and
M.C. Constantinou, 6/15/12.
MCEER-12-0006 Bayesian Fragility for Nonstructural Systems, by C.H. Lee and M.D. Grigoriu, 9/12/12.
MCEER-12-0007 A Numerical Model for Capturing the In-Plane Seismic Response of Interior Metal Stud Partition Walls,
by R.L. Wood and T.C. Hutchinson, 9/12/12.
MCEER-12-0008 Assessment of Floor Accelerations in Yielding Buildings, by J.D. Wieser, G. Pekcan, A.E. Zaghi, A.M.
Itani and E. Maragakis, 10/5/12.
MCEER-13-0001 Experimental Seismic Study of Pressurized Fire Sprinkler Piping Systems, by Y. Tian, A. Filiatrault and
G. Mosqueda, 4/8/13.
MCEER-13-0002 Enhancing Resource Coordination for Multi-Modal Evacuation Planning, by D.B. Hess, B.W. Conley and
C.M. Farrell, 2/8/13.
MCEER-13-0003 Seismic Response of Base Isolated Buildings Considering Pounding to Moat Walls, by A. Masroor and G.
Mosqueda, 2/26/13.
MCEER-13-0004 Seismic Response Control of Structures Using a Novel Adaptive Passive Negative Stiffness Device, by
D.T.R. Pasala, A.A. Sarlis, S. Nagarajaiah, A.M. Reinhorn, M.C. Constantinou and D.P. Taylor, 6/10/13.
MCEER-13-0005 Negative Stiffness Device for Seismic Protection of Structures, by A.A. Sarlis, D.T.R. Pasala, M.C.
Constantinou, A.M. Reinhorn, S. Nagarajaiah and D.P. Taylor, 6/12/13.
MCEER-13-0006 Emilia Earthquake of May 20, 2012 in Northern Italy: Rebuilding a Resilient Community to Withstand
Multiple Hazards, by G.P. Cimellaro, M. Chiriatti, A.M. Reinhorn and L. Tirca, June 30, 2013.
MCEER-13-0007 Precast Concrete Segmental Components and Systems for Accelerated Bridge Construction in Seismic
Regions, by A.J. Aref, G.C. Lee, Y.C. Ou and P. Sideris, with contributions from K.C. Chang, S. Chen, A.
Filiatrault and Y. Zhou, June 13, 2013.
MCEER-13-0008 A Study of U.S. Bridge Failures (1980-2012), by G.C. Lee, S.B. Mohan, C. Huang and B.N. Fard, June 15,
2013.
MCEER-13-0009 Development of a Database Framework for Modeling Damaged Bridges, by G.C. Lee, J.C. Qi and C.
Huang, June 16, 2013.

434

MCEER-13-0010 Model of Triple Friction Pendulum Bearing for General Geometric and Frictional Parameters and for Uplift
Conditions, by A.A. Sarlis and M.C. Constantinou, July 1, 2013.
MCEER-13-0011 Shake Table Testing of Triple Friction Pendulum Isolators under Extreme Conditions, by A.A. Sarlis,
M.C. Constantinou and A.M. Reinhorn, July 2, 2013.
MCEER-13-0012 Theoretical Framework for the Development of MH-LRFD, by G.C. Lee (coordinating author), H.A
Capers, Jr., C. Huang, J.M. Kulicki, Z. Liang, T. Murphy, J.J.D. Shen, M. Shinozuka and P.W.H. Yen, July
31, 2013.
MCEER-13-0013 Seismic Protection of Highway Bridges with Negative Stiffness Devices, by N.K.A. Attary, M.D. Symans,
S. Nagarajaiah, A.M. Reinhorn, M.C. Constantinou, A.A. Sarlis, D.T.R. Pasala, and D.P. Taylor, September
3, 2014.
MCEER-14-0001 Simplified Seismic Collapse Capacity-Based Evaluation and Design of Frame Buildings with and without
Supplemental Damping Systems, by M. Hamidia, A. Filiatrault, and A. Aref, May 19, 2014.
MCEER-14-0002 Comprehensive Analytical Seismic Fragility of Fire Sprinkler Piping Systems, by Siavash Soroushian,
Emmanuel Manos Maragakis, Arash E. Zaghi, Alicia Echevarria, Yuan Tian and Andre Filiatrault, August
26, 2014.
MCEER-14-0003 Hybrid Simulation of the Seismic Response of a Steel Moment Frame Building Structure through
Collapse, by M. Del Carpio Ramos, G. Mosqueda and D.G. Lignos, October 20, 2014.
MCEER-14-0005 Seismic Performance of Steel Plate Shear Walls Considering Various Design Approaches, by R. Purba and
M. Bruneau, October 31, 2014.
MCEER-14-0006 Air-Blast Effects on Civil Structures, by Jinwon Shin, Andrew S. Whittaker, Amjad J. Aref and David
Cormie, October 30, 2014.

435

Characterizing the Rotational Components of Earthquake Ground Motion

University at Buffalo, The State University of New York


133A Ketter Hall Buffalo, New York 14260-4300
Phone: (716) 645-3391 Fax: (716) 645-3399
Email: mceer@buffalo.edu Web: http://mceer.buffalo.edu

ISSN 1520-295X

Characterizing the Rotational


Components of Earthquake
Ground Motion

by
Dhiman Basu, Andrew S. Whittaker
and Michael C. Constantinou

Technical Report MCEER-12-0005


June 15, 2012

ISSN 1520-295X

MCEER-12-0005

This research was conducted at the University at Buffalo, State University of New York and was
supported by MCEER Thrust Area 3, Innovative Technologies.

Anda mungkin juga menyukai